arxiv: v3 [astro-ph.co] 25 May 2015

Size: px
Start display at page:

Download "arxiv: v3 [astro-ph.co] 25 May 2015"

Transcription

1 Mon. Not. R. Astron. Soc. 000, 1 21 (2014) Printed 26 May 2015 (MN LATEX style file v2.2) Baryon effects on the internal structure of ΛCDM halos in the EAGLE simulations arxiv: v3 [astro-ph.co] 25 May 2015 Matthieu Schaller 1, Carlos S. Frenk 1, Richard G. Bower 1, Tom Theuns 1,2, Adrian Jenkins 1, Joop Schaye 3, Robert A. Crain 3,4, Michelle Furlong 1, Claudio Dalla Vecchia 5,6 and I. G. McCarthy 4 1 Institute for Computational Cosmology, Durham University, South Road, Durham, UK, DH1 3LE 2 Department of Physics, University of Antwerp, Campus Groenenborger, Groenenborgerlaan 171, B-2020 Antwerp, Belgium 3 Leiden Observatory, Leiden University, P.O. Box 9513, 2300 RA Leiden, The Netherlands 4 Astrophysics Research Institute, Liverpool John Moores University, 146 Brownlow Hill, Liverpool L3 5RF, UK 5 Instituto de Astrofísica de Canarias, C/ Vía Láctea s/n, La Laguna, Tenerife, Spain 6 Departamento de Astrofísica, Universidad de La Laguna, Av. del Astrofísico Franciso Sánchez s/n, La Laguna, Tenerife, Spain 26 May INTRODUCTION The development of efficient computational techniques and the growing availablity of computing power over the past three decades have made it possible to simulate the evolution of representative cosmological volumes at high enough resolution to follow the formation of cosmic structures over many orders of magnitude in mass. One of the best established and most robust results from this programme is the characterization of the density structure of dark matter (DM) halos in equilibrium whose spherically averaged density profile, ρ(r), is nearly universal in shape and obeys simple scaling relations (Navarro et al. 1996, 1997). The functional form of this NFW radial profile is independent of mass, formation redshift, and cosmological parameters and has the form: matthieu.schaller@durham.ac.uk ABSTRACT We investigate the internal structure and density profiles of halos of mass 0 4 M in the Evolution and Assembly of Galaxies and their Environment (EAGLE) simulations. These follow the formation of galaxies in a ΛCDM Universe and include a treatment of the baryon physics thought to be relevant. The EAGLE simulations reproduce the observed present-day galaxy stellar mass function, as well as many other properties of the galaxy population as a function of time. We find significant differences between the masses of halos in the EAGLE simulations and in simulations that follow only the dark matter component. Nevertheless, halos are well described by the Navarro-Frenk-White (NFW) density profile at radii larger than 5% of the virial radius but, closer to the centre, the presence of stars can produce cuspier profiles. Central enhancements in the total mass profile are most important in halos of mass 2 3 M, where the stellar fraction peaks. Over the radial range where they are well resolved, the resulting galaxy rotation curves are in very good agreement with observational data for galaxies with stellar mass M < 5 0 M. We present an empirical fitting function that describes the total mass profiles and show that its parameters are strongly correlated with halo mass. Key words: cosmology: theory, dark matter, large-scale structure of Universe ρ(r) δ c = ρ cr (r/r s) (1 + r/r s) 2, (1) where ρ cr is the critical density of the Universe, δ c a characteristic density and r s a characteristic radius. Navarro et al. (1997) showed that these two scale parameters are strongly correlated and that the characteristic density is proportional to the density of the universe at the time when the halo was assembled. This proportionality constant or, equivalently, the proportionality constant between halo mass and concentration has been studied by many authors (e.g. Avila-Reese et al. 1999; Jing 0; Bullock et al. 1; Eke et al. 1; Zhao et al. 3; Neto et al. 7; Duffy et al. 8; Gao et al. 8; Navarro et al. 2010; Ludlow et al. 2014; Dutton & Macciò 2014). The validity of the model is well established and a physical understanding of the universality of the profile is beginning to emerge (Ludlow et al. 2013; Correa et al. 2014, 2015). The nearly scale-free behaviour induced by gravity applies only to halos made entirely of DM. In practice, halos of mass above

2 2 M. Schaller et al M participate in the process of galaxy formation. The cooling and dissipation of gas in these halos introduces a characteristic scale that breaks self-similarity (White & Rees 1978; White & Frenk 1991) and the subsequent formation of stars can deepen the potential well and modify the structure of the halo in this region. One of the early models of the effects of baryon collapse on the structure of a halo, making use of adiabatic invariants, concluded that halos would become denser in their centres (Blumenthal et al. 1986). These simple models, however, were later shown not to match hydrodynamic simulations and led to a more general framework for calculating adiabatic contraction based on the average radial distribution of particles (Gnedin et al. 4; Gustafsson et al. 6). The parameters of this model, however, have been shown to depend on halo mass, redshift and on the details of the hydrodynamic simulation, making analytical descriptions of adiabatic contraction complex and uncertain (Duffy et al. 2010). Baryons, however, can also produce the opposite effect and induce expansion rather than contraction of the halo. Using idealized hydrodynamic simulations, Navarro et al. (1996) showed that the rapid expulsion of gas that had previously cooled to very high density near the centre of a halo could generate a central core. Subsequent work using cosmological simulations has confirmed and extended this result (e.g. Read & Gilmore 5; Dehnen 5; Mashchenko et al. 6; Governato et al. 2010; Pontzen & Governato 2012; Teyssier et al. 2013; Martizzi et al. 2013). The structure of the inner halo is often used as a test of the ΛCDM paradigm (e.g. Sand et al. 2; Gilmore et al. 7). Such tests, however, tend to compare observations of halos which have galaxies within them with results from simulations of pure dark matter halos (Newman et al. 2013). For the tests to be meaningful, accurate and reliable calculations of how baryons affect the structure of the halos are essential. Such calculations are also of major importance for efforts to detect DM experimentally, either directly in the laboratory, or indirectly through the products of particle decay or annihilation. Simulating the evolution of the visible components of the universe is a much more complex task than simulating the evolution of the DM because baryons undergo a variety of astrophysical processes many of which are relatively poorly understood. The resolution that is attainable even with the largest computers today is insufficient for an ab initio calculation of most of these processes which, as a result, need to be treated through parametrized subgrid models added to the coupled hydrodynamical and gravitational evolution equations. These models describe the effects of radiative cooling, star formation, feedback from energy liberated during the evolution of stars and supermassive black holes growing at the centres of galaxies. Simulations that include some or all of these processes have shown that significant changes can be induced in the total masses of halos (Sawala et al. 2013, 2015; Cusworth et al. 2014; Velliscig et al. 2014; Vogelsberger et al. 2014) and in their inner structure (e.g. Gnedin et al. 4; Pedrosa et al. 9; Duffy et al. 2010; Pontzen & Governato 2012; Brook et al. 2012; Di Cintio et al. 2014). In this paper we investigate how baryon physics modifies the structure of DM halos in the Evolution and Assembly of Galaxies and their Environment (EAGLE) cosmological hydrodynamical simulations (Schaye et al. 2015). An important advantage of these simulations is that they give a good match to the stellar mass function and and to the distribution of galaxy sizes over a large range of stellar masses (( ) M ). Furthermore, the relatively large volume of the reference EAGLE simulation provides a large statistical sample to derive the halo mass function in the mass range ( ) M and to investigate the radial density profiles of halos more massive than 1 M. This paper is organised as follows. In Section 2 we introduce the simulations and describe the selection of halos. In Section 3 we focus on the change in the mass of halos induced by baryon processes and the effect this has on the halo mass function. In Section 4 we analyse the radial density profile of the halos and decompose them according to their different constituents. We fit the total matter profile with a universal formula that accounts for deviations from the NFW profile and show that the best fit parameters of these fits correlate with the mass of the halo. Our main results are summarized in Section 5. All our results are restricted to redshift z = 0 and all quantities are given in physical units (without factors of h). 2 THE SIMULATIONS The simulations analysed in this paper were run as part of a Virgo Consortium project called the Evolution and Assembly of Galaxies and their Environment (EAGLE; Schaye et al. 2015). The EAGLE project consists of simulations of ΛCDM cosmological volumes with sufficient size and resolution to model the formation and evolution of galaxies of a wide range of masses, and also include a counterpart set of dark matter-only simulations of these volumes. The galaxy formation simulations include the correct proportion of baryons and model gas hydrodynamics and radiative cooling. State-of-the-art subgrid models are used to follow star formation and feedback processes from both stars and AGN. The parameters of the subgrid model have been calibrated to match certain observables as detailed in Schaye et al. (2015). In particular, the simulations reproduce the observed present day stellar mass function, galaxy sizes and many other properties of galaxies and the intergalactic medium remarkably well. These simulations also show the correct trends with redshift of many galaxy properties (Schaye et al. 2015; Furlong et al. 2014). The simulations were run using an extensively modified version of the code GADGET-3 (Springel et al. 8), which is essentially a more computationally efficient version of the public code GADGET-2 described in detail by Springel (5). GADGET uses a Tree-PM method to compute the gravitational forces between the N-body particles and implements the equations of hydrodynamics using Smooth Particle Hydrodynamics (SPH, Monaghan 1992; Price 2010). The EAGLE version of GADGET-3 uses an SPH implementation called ANARCHY (Dalla Vecchia in prep.), which is based on the general formalism described by Hopkins (2013), with improvements to the kernel functions (Dehnen & Aly 2012) and viscosity terms (Cullen & Dehnen 2010). This new implementation of SPH alleviates some of the problems associated with modelling contact discontinuities and fluid instabilities. As discussed by Dalla Vecchia (in prep.), the new formalism improves on the treatment of instabilities associated with cold flows and filaments and on the evolution of the entropy of hot gas in halos. The timestep limiter of Durier & Dalla Vecchia (2012) is applied to ensure good energy conservation everywhere, including regions disturbed by violent feedback due to supernovae and AGN. The impact of this new hydrodynamics scheme on our galaxy formation model is discussed by Schaller et al. (in prep.). The analysis in this paper focusses on two simulations: the Ref-L100N1504 simulation introduced by Schaye et al. (2015), which is the largest EAGLE simulation run to date, and its counterpart dark matter-only simulation, DM-L100N1504. To investigate

3 Baryon effects on the internal structure of ΛCDM halos 3 smaller mass halos and test for convergence in our results we also analyse the higher resolution Recal-L025N0752 simulation (and its dark matter-only counterpart) in which some of the sub-grid physics parameters were adjusted to ensure that this calculation also reproduces the observed galaxy stellar mass function, particularly at the low-mass end, as discussed by (Schaye et al. 2015). We will refer to the two simulations with baryon physics as EAGLE simulations and to the ones involving only dark matter as simulations. The main EAGLE simulation models a cubic volume of sidelength 100 Mpc with gas and dark matter particles to redshift z = 0. A detailed description of the initial conditions is given in Schaye et al. (2015). Briefly, the starting redshift was z = 127; the initial displacements and velocities were calculated using second order Lagrangian perturbation theory with the method of Jenkins (2010); the linear phases were taken from the public multiscale Gaussian white noise field, Panphasia (Jenkins 2013); the cosmological parameters were set to the best fit ΛCDM values given by the Planck-1 data (Planck Collaboration et al. 2014): [Ω m, Ω b, Ω Λ, h, σ 8, n s] = [0.307, , 0.693, , , ]; and the primordial mass fraction of He was set to These choices lead to a dark matter particle mass of M and an initial gas particle mass of M. We use a comoving softening of 2.66 kpc at early times, which freezes at a maximum physical value of 700 pc at z = 2.8. The Recal-L025N0752 simulation follows gas and DM particles in a 25 Mpc volume assuming the same cosmological parameters. This implies a DM particle mass of M and an initial gas mass of M. The softening is 1.33 kpc initially and reaches a maximum physical size of 350 pc at z = 0. The simulations, DM-L100N1504 and DM-L025N0752, follow exactly the same volume as EAGLE, but with only and collisionless dark matter particles, each of mass M and M, respectively. All other cosmological and numerical parameters are the same as in the EAGLE simulation. 2.1 Baryonic physics The baryon physics in our simulation correspond to the Ref EA- GLE model. The model, fully described in Schaye et al. (2015), is summarized here for completeness. Star formation is implemented following Schaye & Dalla Vecchia (8). A polytropic equation of state, P ρ 4/3, sets a lower limit to the gas pressure. The star formation rate per unit mass of these particles is computed using the gas pressure using an analytical formula designed to reproduce the observed Kennicutt-Schmidt law (Kennicutt 1998) in disk galaxies (Schaye & Dalla Vecchia 8). Gas particles are converted into stars stochastically. The threshold in hydrogen density required to form stars is metallicity dependent with lower metallicity gas having a higher threshold, thus capturing the metallicity dependence of the HI H 2 phase transition (Schaye 4). The stellar initial mass function is assumed to be that of Chabrier (3) in the range 0.1M to 100M with each particle representing a single age stellar population. After yrs all stars with an initial mass above 6M are assumed to explode as supernovae. The energy from these explosions is transferred as heat to the surrounding gas. The temperature of an appropriate amount of surrounding gas is raised instantly by K. This heating is implemented stochastically on one or more gas particles in the neighbourhood of the explosion site (Dalla Vecchia & Schaye 2012). This gas, once heated, remains coupled in a hydrodynamic sense with its SPH neighbours in the ISM, and therefore exerts a form of feedback locally that can affect future star formation and radiative cooling. The energy injected into the gas corresponds to erg per supernovae times a dimensionless efficiency factor, f E, that depends on the local gas metallicity and density. The construction of f E and its impact on galaxy formation is discussed thoroughly by Schaye et al. (2015) and Crain et al. (2015). For a gas of metallicity, Z, and hydrogen number density, n H, the efficiency in the reference model is: f E = S (X; w), (2) where w = 2/ ln 10, ( ) ( ) Z 0.1 cm 3 X = 3.35, (3) 0.1Z n H and S(X; w) is a convenient sigmoid function which varies between 0 and 1, and which we will need again in the following section. We define the sigmoid function for x 0, w > 0 as S(X; w) = Xw 1 + X w. (4) As X varies from zero to infinity, the sigmoid function S(X; w) smoothly varies between 0 and 1, taking the value of 1 when the 2 argument X = 1. The parameter w controls the rapidity of the transition between the asymptotes. Besides energy from star formation, the star particles also release metals into the ISM through three evolutionary channels: type Ia supernovae, winds and supernovae from massive stars, and AGB stars using the method discussed in Wiersma et al. (9b). The yields for each process are taken from Portinari et al. (1998), Marigo (1) and Thielemann et al. (3). Following Wiersma et al. (9a), the abundances of the eleven elements that dominate the cooling rates are tracked. These are used to compute elementby-element dependent cooling rates in the presence of the Cosmic Microwave Background and the ultraviolet and X-ray backgrounds from galaxies and quasars according to the model of Haardt & Madau (1). For halos whose masses first exceed M FOF = 0 h 1 M ( 1500 dark matter particles, see section 2.2), black hole (BH) sink particles are placed at the centre of the halos. The BHs are then allowed to grow through gas accretion and by merging with other BHs using methods based on those introduced by Springel et al. (5) and Booth & Schaye (9). The gas surrounding a BH is accreted at a rate given by the Bondi-Hoyle formula (Bondi & Hoyle 1944) unless the viscous timescale of the gas around the BH is larger than the Bondi time, in which case the accretion rate is reduced by a factor proportional to the cube of the ratio of the local sound speed and the rotation velocity (Rosas-Guevara et al. 2013). For a BH of mass, M BH, surrounded by gas at density, ρ, velocity with respect to the BH, v, and sound speed, c s, the accretion rate is: { ṁ BH = 4πGM ( ) 3 BHρ 2 (c 2 s + v 2 ) 1 c s 3/2 C visc V φ if C viscvφ 3 > c 3 s, (5) 1 if C viscvφ 3 c 3 s where V φ is the circular speed of the gas at the Bondi radius and C visc = 2π in the reference simulation. Feedback due to AGN activity is implemented in a similar way to the feedback from star formation described above. The fraction of the accreted rest mass energy liberated by accretion is ɛ r = 0.1, and the heating efficiency of this liberated energy (i.e. the fraction

4 4 M. Schaller et al. of the energy that couples to the gas phase) is ɛ f = Gas particles receiving AGN feedback energy are chosen stochastically and their temperature is raised by K. These models of supernova and AGN feedback are extensions of the models developed for the Virgo Consortium projects OWLS (Schaye et al. 2010) and GIMIC (Crain et al. 9). The values of the parameters were constrained by matching key observables of the galaxy population including the observed z 0 galaxy stellar mass function, galaxy sizes and the relation between black hole and stellar mass (Crain et al. 2015). 2.2 Halo definition and selection Halos were identified using the Friends-of-Friends (FOF) algorithm on all dark matter particles adopting a dimensionless linking length, b = 0.2 (Davis et al. 1985). We then applied the SUBFIND algorithm, which is built into GADGET-3 (Springel et al. 1; Dolag et al. 9), to split the FOF groups into self-bound substructures. A sphere is grown outwards from the potential minimum of the dominant subgroup out to a radius where the mean interior density equals a target value. This target value is conventionally defined in units of the critical density, ρ cr(z) = 3H 2 (z)/8πg. With our choice of cosmology, at z = 0 we have ρ cr = ρ cr(0) = M kpc 3. A halo of mass, M X, is then defined as all the mass within the radius, R X, for which 3M X = Xρ cr(z) (6) 4πRX 3 Commonly used values are X =, 500 and 2500, leading to the definition of the mass, M, and the radius, R, and similar definitions for other values of X. In the particular case of the virial radius, R vir, one can use the spherical top-hat collapse model to derive the value of X (Eke et al. 1996). We use the fitting formula given by Bryan & Norman (1998): X = 18π (Ω m(z) 1) 39 (Ω m(z) 1) 2, (7) where ( ) 2 Ω m(z) = Ω m (1 + z) 3 H0, (8) H(z) and H(z) is the value of the Hubble parameter at redshift z which, in a flat Universe, is H(z) = H 0 Ωm(1 + z) 3 + Ω Λ. (9) In the case of the Planck1 cosmology, at z = 0, X = 102.1, giving M vir = M 102 and R vir = R 102. We define the circular velocity, V X, as GMX V X =. (10) R X We only consider halos with more than particles within R, implying a limit, M M, in our joint analysis of the two EAGLE simulations. For specific properties that depend on the internal structure of the halo we adopt more conservative limits as described in section Matching halos between the two simulations The EAGLE and simulations start from identical Gaussian density fluctuations. Even at z = 0 it is possible, in most cases, to identify matches between halos in the two simulations. These matched halos are comprised of matter that originates from the same spatial locations at high redshift in the two simulations. In practice, these identifications are made by matching the particle IDs in the two simulations, as the values of the IDs encode the Lagrangian coordinates of the particles in the same way in both simulations. For every FOF group in the EAGLE simulation, we select the 50 most bound dark matter particles. We then locate those particles in the simulation. If more than half of them are found in a single FOF group in the simulation, we make a link between those two halos. We then repeat the procedure by looping over FOF groups in the simulation and looking for the position of their counterparts in the EAGLE simulation. More than 95% of the halos with M > 2 0 M can be matched bijectively, with the fraction reaching unity for halos above 7 0 M in the L100N1504 volumes. Similarly, 95% of the halos with M > can be matched bijectively in the L025N0752 volumes. 3 HALO MASSES AND CONTENT Previous work comparing the masses of halos in cosmological galaxy formation simulations with matched halos in counterpart dark matter-only simulations have found strong effects for all but the most massive halos (e.g. Cui et al. 2012; Sawala et al. 2013). Sawala et al. (2013) found that baryonic effects can reduce the masses of halos by up to 25% for halo masses (in the dark matter only simulation) below 3 M. (They did not include AGN feedback in their simulation.) A similar trend was observed at even higher masses by Martizzi et al. (2013), Velliscig et al. (2014), Cui et al. (2014) and Cusworth et al. (2014) using a variety of subgrid models for star formation and stellar and AGN feedback. All these authors stress that their results depend on the details of the subgrid implementation used. This is most clearly shown in Velliscig et al. (2014), where the amplitude of this shift in mass is shown explicitly to depend on the subgrid AGN feedback heating temperature, for example. Hence, it is important to use simulations that have been calibrated to reproduce the observed stellar mass function. In this section we find that similar differences to those seen before occur between halo masses in the EAGLE and models. These differences are of particular interest because EAGLE reproduces well a range of low-redshift observables of the galaxy population such as masses, sizes and star formation rates (Schaye et al. 2015), although the properties of clusters of galaxies are not reproduced as well as in the Cosmo-OWLS simulation (Le Brun et al. 2014) analyzed by Velliscig et al. (2014). 3.1 The effect of baryon physics on the total halo mass In this section we compare the masses of halos in the EAGLE and simulations combining our simulations at two different resolutions. To minimise any possible biases due to incomplete matching between the simulations, we only consider halos above M (in ), since these can be matched bijectively to their counterparts in more than 95% of cases. Fig. 1 shows the ratio of M for matched halos in the EA- GLE and simulations as a function of M in the simulation. The black filled circles correspond to the geometric mean of the ratios in each logarithmically spaced mass bin. The choice of a geometric mean is motivated simply by the fact that its reciprocal is

5 Baryon effects on the internal structure of ΛCDM halos z = 0 Table 1. Best fitting parameters to the black points in Fig. 1 using Eqn. 13, and their uncertainties which are taken to be the diagonal elements of the correlation matrix of the least-squares fitting procedure. M EAGLE /M EAGLE M 12 = 1.3 M M 23 = 3.2 M M [M ] f univ DM = Ωm Ω b Ω m = Figure 1. The ratio of the masses of the matched halos in the EAGLE and simulations. The red squares show values for individual halos and the black filled circles values binned by halo mass. Halos with M < 0.1 M are extracted from the higher resolution, L025N0752, simulation. The binned points are the geometric average of the individual ratios with the error bars at M < 0.1 M indicating the uncertainty arising from the low number of halos in the high-resolution simulation. The black dashed lines placed above and below the black points show the geometrical 1σ scatter for each bin. The lower horizontal grey dotted line indicates the universal dark matter fraction f DM = 1 f b = (Ω m Ω b )/Ω m = The upper dotted line marks unity. The green solid line is the function of Eqn. 13 fitted to the binned ratios. The vertical dotted lines mark the values of the fitting parameters M 12 and M 23. the geometric mean of M /M EAGLE, which is also a quantity of interest. The halos in EAGLE are typically lighter than their counterparts. There appear to be three distinct regimes in Fig. 1. At the low mass end, M < 5 0 M, M EAGLE /M drops to This is less than one minus the universal baryon fraction, f DM, so not only have the baryons been removed but the dark matter has also been disturbed. The reduction in mass due to the loss of baryons lowers the value of R and thus the value of M. However, this reduction in radius is not the sole cause for the reduction in halo mass: the amount of mass within a fixed physical radius is actually lower in the simulation with baryons because the loss of baryons, which occurs early on, reduces the growth rate of the halo (Sawala et al. 2013). At higher masses, stellar feedback becomes less effective, but AGN feedback can still expel baryons and the ratio rises to a plateau of 0.85 between M = 2 M and 5 2 M. Finally, for the most massive halos (M > 4 M ) not even AGN feedback can eject significant amounts of baryons from the halos and the mass ratio asymptotes to unity. Sawala et al. (2013) proposed a fitting function to the ratio of M in simulations with and without baryons from the GIMIC project (Crain et al. 9). Their study focused mostly on lowermass objects and subhalos, but included enough large halos to sample the high-mass end of the relation. Their four parameter fitting function can be written as: M = a + (b a)s M ( M ; w M t ), (11) where S is a sigmoid function that varies smoothly between 0 and 1, and is defined in Eqn. 4. The best-fit parameter values in Sawala et al. (2013) are: (a, b, log 10 (M t/m ), w) = Parameter Value 1 σ fit uncertainty r ± r ± r ± log 10 (M 12 /M ) ±0.003 log 10 (M 23 /M ) ±0.029 t ±0.045 t ±0.18 (0.69, 0.98, 11.6, 0.79). The values of a and b correspond to the low- and high-mass asymptotes, respectively. Velliscig et al. (2014) used a similar fitting function to summarise the results of their study, again with four parameters, which can be written as: M = a M ( b a ) S(M /M t ;w), (12) where exactly the same sigmoid function is used to interpolate between the two asymptotic values, a and b, but now in a geometric rather than arithmetic fashion. The functional forms of Eqns. 11 and 12 are virtually identical as, in practice, the ratio b/a is never far from unity. It is quite clear, however, from Fig. 1 that a single sigmoid function does not reproduce the behaviour we observe particularly well: the ratio shows three, not two, distinct plateaux. The simulations used by Sawala et al. (2013) did not include AGN feedback and so did not show the change in mass arising from this form of feedback. In contrast, the simulations used by Velliscig et al. (2014) did not have sufficient numerical resolution to see the asymptotic low-mass behaviour determined by stellar feedback. To fit our results, we use a double sigmoid: M = r M 1 + (r 2 r 1)S + (r 3 r 2)S ( M ) ; t 12 M 12 ), (13) ( M M 23 ; t 23 where the seven parameters can be interpreted as follows: r 1, r 2 and r 3 are the values of the ratios corresponding to the three distinct plateaux; the mass scales, M 12 and M 23, are the mid-points between regimes 1 and 2, and 2 and 3 respectively; and the parameters, t 12 and t 23, control the rapidity of each transition. The green curve in Fig. 1 shows the best fitting curve to the black binned data points. The fit was obtained by a least-squares minimisation for all seven parameters assuming Poisson uncertainties for each mass bin. Adopting a constant error instead gives very similar values for all parameters. The values of the two transition masses, M 12 and M 23, are shown as vertical dotted lines in Fig. 1. The best-fitting parameters are given in Table 1. Note that the value of r 3 is, as expected, very close to unity. The value of the first transition mass, M 12 = 1.35 M, is similar to that reported by Sawala et al. (2013) who found M t = 1.6 M for the GIMIC simulations. The second transition, M 32 = 3.2 M, is located well below the range of values found by Velliscig et al. (2014) (3.7 M M ). However, as Schaye et al. (2015) have shown the AGN feedback in the few rich clusters formed in the EAGLE volume may not be strong enough,

6 6 M. Schaller et al. dn/d(log 10 (M)) [Mpc 3 ] Ratio NLY particles (L025N0752) NLY particles (L100N1504) M 0.9 NLY EAGLE M [M ] Figure 2. Top panel: the abundance of halos at z = 0 as a function of the mass, M, in the EAGLE (red curve, lower line) and (green curve, upper line) simulations. The high resolution volume is used for M < 0.1 M. The resolution limits for both simulations are indicated by the vertical dashed lines on the left, and the number of halos in sparsely populated bins is given above the Poisson error bars. Bottom panel: the ratio of the mass functions in the EAGLE and simulations. as evidenced by the fact that this simulation overestimates the gas fractions in clusters, whereas the 400 Mpc/h Cosmo-OWLS simulation used by Velliscig et al. (2014) reproduces these observations (Le Brun et al. 2014). A simulation with stronger AGN feedback, EAGLE-AGNdT9, which gives a better match to the group gas fractions and X-ray luminosities than EAGLE, was discussed by Schaye et al. (2015). Applying the same halo matching procedure to this simulation and its collisionless dark matter-only counterpart, we obtain slightly different values for the best-fitting parameters of Eqn. 13. The difference is mainly in the parameters, M 23 and t 23, which describe the high-mass end of the double-sigmoid function. In this model, the transition occurs at log 10 (M 23/M ) = ± 0.09, closer to the values found by Velliscig et al. (2014). The width of the transition, however, is poorly constrained, t 23 = 3.0 ± 12.7, due to the small number of halos (only eight with M, > 2 3 M ) in this simulation which had only an eighth the volume of the reference simulation. As Velliscig et al. (2014) did, we provide a fit to the scatter in the log of the ratio about the mean relation, valid over the range where appropriately constraining data are available: ( ) ( ) σ log 10 (M M ) = log 10. (14) 2 M The scatter is about 10% for a halo mass of 2 M and decreases with mass. The slope in the relation is approximatively a factor of two greater than that found for the AGN models of Velliscig et al. (2014). 3.2 The halo mass function The effect of baryons on the halo mass function can be seen in Fig. 2. The red and green lines in the top panel show the mass functions in the EAGLE and simulations. The ratio of the two functions (bottom panel) shows an almost constant shift over most of the plotted mass range, M /M = , as expected from Fig. 1. The relatively small volume of the EAGLE simulation does not sample the knee of the halo mass function well, but extrapolating the fit to the mass ratios of Eqn. 13 to higher masses, together with results from previous studies (Cusworth et al. 2014; Martizzi et al. 2013; Velliscig et al. 2014), suggests that the differences vanish for the most massive objects. Studies that rely on galaxy clusters to infer cosmological parameters will need to take account of the effects of the baryons, particularly for clusters of mass M 4 M. 3.3 Baryonic and stellar fractions in the EAGLE simulation We have shown in the previous subsection that for all but the most massive examples, halo masses are systematically lower when baryonic processes are included. In this subsection we examine the baryonic content of halos in the EAGLE simulation. We restrict our analysis to the L100N1504 volume. Fig. 3 shows the mass fractions of baryons and stars within R as a function of the halo mass, M, in the EAGLE simulation. The baryon fraction increases with halo mass and approaches the universal mean value, fb univ Ω b /Ω m, for cluster mass halos. The gas is the most important baryonic component in terms of mass over the entire halo mass range. At a much lower amplitude everywhere, the stellar mass fraction peaks around a halo mass scale of 2 2 M where star formation is at its least inefficient. The baryon fractions are much lower than the universal value for all but the most massive halos. For Milky Way sized halos, we find f b /fb univ It is only for group and cluster sized halos, whose deeper gravitational potentials are able to retain most of the baryons even in the presence of powerful AGN, that the baryon fraction is close to fb univ. The baryon fractions of the halos extracted from the EAGLE-AGNdT9 model (which provides a better match to X-ray luminosities; Schaye et al. 2015) are presented in Appendix A1. The stellar mass fraction is never more than a few percent. At the peak, around M 2 2 M, it reaches a value of Multiplying the stellar fraction by the halo mass function leads to an approximate stellar mass function, which is close to the actual one (published in Schaye et al. 2015), after a fixed aperture correction is applied to mimic observational measurements. As may be seen in both panels, there is significant scatter in the baryonic and stellar fractions, with variations of a factor of a few possible for individual halos. While the baryonic and stellar fractions are low within R, they are much higher in the inner regions of halos as shown in Fig. 4, where these fractions are now plotted within 0.05R, a scale commensurate with the sizes of galaxies both in EAGLE and in the real universe. Within this radius the fractions rise above the cosmic mean for halos in the mass range 5 1 M < M < 2 3 M. The central parts of these halos are strongly dominated by the baryons. In agreement with observations of the nearby universe, the most important contribution to the mass on these scales is from stars rather than gas. Another notable feature is that the most massive halos are baryon poor in their central regions, reflecting the regulation by AGN feedback.

7 Baryon effects on the internal structure of ΛCDM halos 7 fb(r < R) f (r < R) M [M ] fb(r < R)/ f univ b f (r < R)/ f univ Figure 3. Baryon fraction, f b = M b /M (top panel), and stellar fraction, f = M /M (bottom panel), within R as a function of M. The right-hand axis gives the fractions in units of the universal mean value, fb univ = The solid circles in the top panel and the stars in the bottom panel show the mean value of the fractions binned by mass. The dashed lines above and below these symbols show the RMS width of each bin with more than three objects. The stellar fractions are reproduced as grey stars in the top panel. fb(r < 0.05 R) f (r < 0.05 R) M [M ] b fb(r < 0.05 R)/ f univ b f (r < 0.05 R)/ f univ Figure 4. Same as Fig. 3 but for the mass contained within 5% of R. Note the different scale on the ordinate axis. The dotted horizontal lines mark one and two times the universal baryon fraction. b 4 HALO PROFILES In this section we explore the effects of baryons on halo profiles restricting the analysis to halos with more than 5000 particles within R vir, which corresponds to a halo mass of about 5 0 M in the L100N1504 simulation and M in the L050N0752 simulation. The stellar masses found in the EAGLE simulation for halos of this mass are consistent with observational expectations based on abundance matching (Schaye et al. 2015). Halos smaller than this typically have fewer than the hundred star particles, which Schaye et al. (2015) showed to be a necessary criterion for many applications. This limit of 5000 in the number of particles is intermediate between those used in other studies. It is similar to the number adopted by Ludlow et al. (2013) and lower than the number adopted by Neto et al. (7) and Duffy et al. (8, 2010) (10000 particles), but higher than the number adopted by Gao et al. (8); Dutton & Macciò (2014) (3000 particles) or Macciò et al. (7) (250 particles). There are halos with at least 5000 particles in the Ref-L100N1504 EAGLE simulation and 2460 in the Recal- L025N0752 simulation. We define relaxed halos as those where the separation between the centre of the potential and the centre of mass is less than 0.07R vir, as proposed by Macciò et al. (7). Neto et al. (7) used this criterion, and also imposed limits on the substructure abundance and virial ratio. Neto et al. (7) found that the first criterion was responsible for rejecting the vast majority of unrelaxed halos. Their next most discriminating criterion was the amount of mass in substructures. In common with Gao et al. (8), here we use stacked profiles. Hence, individual substructures, which can be important when fitting individual halos, have a smaller effect on the average profile. We therefore do not use a substructure criterion to reject halos. Our relaxed sample includes halos in the L100N1504 simulation and 1590 in the L025N0752 simulation. We construct the stacked halos by coadding halos in a set of contiguous bins of width log 10 (M ) = 0.2. The density and mass profiles of each halo and of the stacked halos are obtained using the procedure described by Neto et al. (7). We define a set of concentric contiguous logarithmically spaced spherical shells of width log 10 (r) = 0.078, with the outermost bin touching the virial radius, R vir. The sum of the masses of the particles in each bin is then computed for each component (dark matter, gas, stars, black holes) and the density is obtained by dividing each sum by the volume of the shell. 4.1 Resolution and convergence considerations Determining the minimum radius above which the results are robust and reliable is non-trivial. For DM-only simulations, Gao et al. (8) showed that the best fit NFW profiles are sensitive to this choice and it is, therefore, important to estimate this minimum converged radius accurately. For DM-only simulations the thorough resolution study of (Power et al. 3, P03) suggests a convergence radius, R P 03, based on the two-body relaxation timescale of particles orbiting in the gravitational potential well. This criterion can be written as: 4πρcr N(< RP 03) ln N(< R P 03) R3/2 P 03, (15) 3m DM where N(< r) is the number of particles of mass, m DM, within radius r. While this criterion could be applied to the simulation, the situation for the EAGLE simulation is more complex since, as

8 8 M. Schaller et al. discussed by Schaye et al. (2015), the concept of numerical convergence for the adopted subgrid model is itself ill defined. One option would be simply to apply the P03 criterion, which is appropriate for the simulation, to both simulations. Alternatively, we could apply the criterion to the dark matter component of the halos in the baryon simulation or to all the collisionless species (stars, dark matter and black holes). Neither of these options is fully satisfactory but, in practice, they lead to similar estimates for R P 03. For the smallest halos of the L100N1504 simulation considered in this section, we find R P kpc whereas for the largest clusters we obtain R P kpc. The original P03 criterion ensures that the mean density internal to the convergence radius, ρ = 3M(r < R P 03)/4πR 3 P 03, is within 10% of the converged value obtained in a simulation of much higher resolution. As the magnitude of the differences between the EAGLE and profiles that we see are significantly larger than 10% typically, we can relax the P03 criterion somewhat. Reanalysing their data, we set the coefficient on the left-hand side of Eqn. 15 to 0.33, which ensures a converged value of the mean interior density at the 20% level. With this definition, our minimal convergence radius r c takes values between 4 kpc and 2.9 kpc for halos with M 1 M up to M 4 M. Similarly, in the L025N0752 simulation our modified criterion gives r c 1.8 kpc. Note that despite adopting a less conservative criterion than P03, the values of r c are always greater than the Plummer equivalent softening length where the force law becomes Newtonian, 2.8ɛ = 0.7 kpc in the L100N1504 simulation and 0.35 kpc in L025N0752 simulation. The validity of our adopted convergence criterion can be tested directly by comparing results from our simulations at two different resolutions. Specifically, we compare our two simulations of (25 Mpc) 3 volumes, L025N0752, and L025N0376 which has the same initial phases as L025N0752 but the resolution of the reference, L100N1504, simulation. In the language of Schaye et al. (2015), this is a weak convergence test since the parameters of the subgrid models have been recalibrated when increasing the resolution. Fig. 5 shows the stacked profiles of the 44 relaxed halos of mass 1 M present in both the L025N0376 and L025N0752 simulations. This mass bin contains enough halos for the stacks not to be dominated by Poisson noise and the halos are large enough to contain more than 5000 particles in the lower resolution simulation. The three panels show density, contained mass and circular velocity profiles respectively, using symbols for the default resolution and lines for the higher resolution simulation. As may be seen, the stacked dark matter and total matter profiles are very well converged over most of the radial range, both in terms of the integral quantities, M(r) and V c(r), and in terms of the differential quantity, ρ(r). The dashed and dotted vertical lines show the convergence radius, r c, for the default and high resolution simulations respectively, computed following the procedure described above. The dark matter and total matter profiles converge well down to much smaller radii than r c implying that this limit is very conservative. This is a consequence of comparing stacked rather than individual halos since the stacks tend to average deviations arising from the additional mass scales represented in the high resolution simulation. We conclude from this analysis that the total matter and dark matter profiles of stacked halos are well converged in our simulations and that we can draw robust conclusions about their properties for r > r c in both the L100N1504 and L025N0752 simulations. The gas profiles in these simulations display a much poorer level of convergence. The disagreement between the two simulations increases at radii larger than r > r c. However, since the mass in gas is negligible at all radii and at all halo masses, the poor convergence of the gas profiles does not affect our conclusions regarding the dark and total matter profiles. We defer the question of the convergence of gaseous profiles to future studies and simulations. 4.2 Stacked halo density and cumulative mass of relaxed halos Having established a robust convergence criterion for stacked halos we now analyse their profiles extracting halos of mass M 1 M from the L100N1504 simulation and halos of mass 0 M M 1 M from the L025N0376 simulation. Fig. 6 shows the stacked profiles for five different halo mass bins. The left-hand column shows that the DM is the dominant component of the density of halos of all masses outside about one percent of R. Inside this radius the stellar component begins to contribute and even dominate in the case of halos with mass 2 M. Considering only the baryonic matter, the inner radii are dominated by stars, but gas dominates outside of 0.1R, as we already saw in Fig. 3. In halos of Milky Way size (M 2 M ) the density profile of the gas is roughly isothermal with ρ(r). The stars exhibit a steep profile, ρ(r) r 4, in the region where this is resolved (r > r c). The resolution of our simulations is not sufficient to enable the discussion of the stellar profile in the central part of the galaxies, within 3 kpc of the centre of potential. The shape of the dark matter profiles in the EAGLE simulation are typically very close to those obtained in the simulation. The profiles depart from the shape in halos with M 2 M, where the slope in the inner regions (below 0.1R ) is slightly steeper. This indicates that some contraction of the dark matter has taken place, presumably induced by the presence of baryons in the central region. The total density profiles of the EAGLE halos also closely resemble those of the simulation. This follows because the DM dominates over the baryons at almost all radii. In halos with a significant stellar fraction, the total profile is dominated by the stars within 0.01R. This creates a total inner profile that is steeper than in the simulations. The stellar contribution is dominant only in the first few kiloparsecs almost independently of the halo mass. Given that halos have profiles similar to an NFW profile, this implies that the total profile will be closer to an NFW for more massive halos because the stars will only be important inside a smaller fraction of the virial radius. This is most clearly seen in the 4 M halo where the profile is dominated by the DM and follows the NFW form down to 0.01R. Similarly, in the smallest halos, M 0 M, the baryon content is so low that the total matter profile behaves almost exactly like the dark matter profile and is hence in very good agreement with dark matter-only simulations. It is also interesting to note the absence in our simulations of DM cores of size kpc such as have been claimed in simulations of individual halos of various masses, assuming different subgrid models and, in some cases, different techniques for solving the hydrodynamical equations (e.g. Navarro et al. 1996; Read & Gilmore 5; Mashchenko et al. 6; Pontzen & Governato 2012; Teyssier et al. 2013; Martizzi et al. 2013; Arraki et al. 2014; Pontzen & Governato 2014; Trujillo-Gomez et al. 2015; Murante et al. 2015; Oñorbe et al. 2015), even though such cores would

9 ρ(r) [M kpc 3 ] rc = 1.7 kpc rc = 3.9 kpc M = 1 M r/r M(< r) [M ] Baryon effects on the internal structure of ΛCDM halos 9 rc = 1.7 kpc rc = 3.9 kpc R = 97 kpc r/r Vc(r) [km s 1 ] rc = 1.7 kpc rc = 3.9 kpc Gas Stars DM Total r [kpc] Figure 5. From left to right: the density, mass and circular velocity profiles of a stack of the 44 relaxed halos of mass 1 M at z = 0 that are present in both the L025N0752 simulation (lines) and the L025N0376 simulation (symbols). Profiles of total matter (green), dark matter (black), gas (blue) and the stellar component (red) are shown for both resolutions. The vertical dashed and dotted lines show the resolution limits, r c, derived from our modified P03 criterion for the L025N0376 and L025N0752 simulations respectively; data point are only shown at radii larger than the Plummer equivalent force softening. The dark matter, total matter and stellar profiles are well converged even at radii smaller than r c, indicating that this convergence cirterion is very conservative when relaxed halos in a narrow mass range are averaged together. Convergence is much poorer for the subdominant gas distribution at large radii. have been resolved in our highest resolution simulations. As first shown by Navarro et al. (1996), density cores can be generated by explosive events in the central regions of halos when gas has become self-gravitating. Our simulations include violent feedback processes but these are not strong enough to generate a core or even a systematic flattening of the inner DM profile on resolved scales. We cannot, of course, rule out the possibility that the central profile could be modified even with our assumed subgrid model in higher resolution simulations. 4.3 Halo circular velocities The right-hand column of Fig. 6 shows the rotation curves. Those for Milky Way mass halos display a flat profile at radii greater than 10 kpc as observed in our galaxy and others (e.g. Reyes et al. 2011). The dominant contribution of the DM is clearly seen here. The stellar component affects only the first few kiloparsecs of the rotation curve. The rotation curves of halos with a significant (> 0.01) stellar fraction (i.e. halos with M > 3 1 M ) have a higher amplitude than the corresponding stacked curves at small radii r 10 kpc. The combination of the stellar component and contraction of the inner dark matter halo leads to a maximum rotation speed that is 30% higher in the EAGLE simulation compared to that in. To assess whether the circular velocity profiles for the galaxies in the EAGLE simulation are realistic, we compare them to a sample of observed disc galaxies. We use the data from Reyes et al. (2011), who observed a sample of 189 spiral galaxies and used Hα lines to measure the circular speeds. From their SDSS r band magnitudes and g r colours, we derive the stellar masses of their galaxies using the M /L scaling relation of Bell et al. (3). We apply a 0.1 dex correction to adjust these stellar mass estimates from their assumed diet Salpeter IMF to our adopted Chabrier (3) IMF, and apply the correction from Dutton et al. (2011) to convert our masses to the MPA/JHU definitions (See McCarthy et al. (2012) for the details.). In Fig. 7 we show the rotation curves of our sample of relaxed halos binned by the stellar mass contained within an aperture of 30 kpc, as used by Schaye et al. (2015) who already compared the predicted maximum circular velocities to observations. The simulated galaxies match the observations exceptionally well, both in terms of the shape and the normalisation of the curves. For all mass bins up to M < 1 M, the EAGLE galaxies lie well within the scatter in the data. Both the shape and the amplitude of the rotation curves are reproduced in the simulation. The scatter appears to be larger in the real than in the simulated population, particularly in the range 10.5 < log 10 M /M < (lower left panel), but the outliers in the data might affected by systematic errors (Reyes et al. 2011) arising, for instance, from the exact position of the slit used to measure spectral features or from orientation uncertainties. The rotation curves for the highest stellar mass bin in the simulation, M > 1 M, show a clear discrepancy with the data. Although the general shape of the curves is still consistent, the normalisation is too high. Part of this discrepancy might be due to the selection of objects entering into this mass bin. The data refer to spiral galaxies, whereas no selection besides stellar mass has been applied to the sample of simulated halos. This highest mass bin is dominated by elliptical objects in EAGLE. Selecting spiral-like objects (in a larger simulation) may well change the results at these high stellar masses. A more careful measurement of the rotation velocities in the simulations in a way that is closer to observational estimates (e.g. by performing mock observations of stellar emission lines) might also reduce the discrepancies. We defer this, more careful, comparison to future work. At all masses beyond the convergence radius the dominant contribution to the rotation curve comes from the dark matter. For the highest mass bins the stellar contribution is very important near the centre and this is crucial in making the galaxy rotation curves relatively flat. As already seen in the previous figure, the contribution of gas is negligible. 4.4 An empirical universal density profile It is well known that the density profiles of relaxed halos extracted from dark matter only simulations are well fit by the NFW profile (Eqn. 1) at all redshifts down to a few percent of the virial radius (Navarro et al. 1997; Bullock et al. 1; Eke et al. 1; Navarro et al. 4; Shaw et al. 6; Macciò et al. 7; Neto et al. 7; Duffy et al. 8; Ludlow et al. 2013; Dutton & Macciò 2014). The total matter profiles shown in Fig. 6 for the EAGLE simulation

10 10 M. Schaller et al. r 2 ρ(r)/(r 2 ρ cr) 10 0 rc = 2.0 kpc M = 0 M M(< r) [M ]1010 rc = 2.0 kpc R = 45 kpc N halo = 362 z = Vc(r) [km s 1 ] rc = 2.0 kpc Gas Stars DM Total NLY r 2 ρ(r)/(r 2 ρ cr) r 2 ρ(r)/(r 2 ρ cr) r 2 ρ(r)/(r 2 ρ cr) r 2 ρ(r)/(r 2 ρ cr) rc = 2.9 kpc rc = 3.2 kpc rc = 3.9 kpc M = 1 M M = 2 M M = 3 M M = 4 M r/r M(< r) [M ] M(< r) [M ] M(< r) [M ] M(< r) [M ] rc = 2.9 kpc rc = 3.2 kpc rc = 3.9 kpc R = 97 kpc N halo = 2412 z = 0.0 R = 209 kpc N halo = 282 z = 0.0 R = 455 kpc N halo = 29 z = 0.0 R = 965 kpc N halo = 1 z = 0.0 r/r Vc(r) [km s 1 ] Vc(r) [km s 1 ] Vc(r) [km s 1 ] Vc(r) [km s 1 ] rc = 3.9 kpc rc = 3.2 kpc rc = 2.9 kpc rc = 2.7 kpc r [kpc] Figure 6. From left to right: the density, mass and circular velocity profiles for stacks of relaxed halos in different mass bins at z = 0. From top to bottom: bins centred on M 0 M, 1 M, 2 M, 3 M and 4 M. Profiles of the total matter (green diamonds), dark matter (black squares), gas (blue circles) and stellar component (red stars) are shown for the halos extracted from the EAGLE simulation. Profiles extracted from halos of similar mass in the simulation are shown with a magenta solid line on all panels. The RMS scatter of the total profile is shown as a green shaded region. The vertical dashed line shows the (conservative) resolution limit, r c, introduced in the previous subsection; data are only shown at radii larger than the force softening. The number of halos in each mass bin is indicated in the middle panel of each row. The density profiles have been multiplied by r 2 and normalized to reduce the dynamic range of the plot and to enable easier comparisons between different halo masses. Note that following the analysis of Section 3.1, matched halos are not guaranteed to fall into the same mass bin. The oscillations seen in the profiles of the two highest mass bins, which have only a few examples, are due to the object-to-object scatter and the presence of substructures.

11 Baryon effects on the internal structure of ΛCDM halos 11 Vc(r) [km s 1 ] rc = 3.7 kpc 9.00 log 10 M /M < 9.25 N gal = 1144 rc = 3.6 kpc 9.25 log 10 M /M < 9.50 N gal = 955 rc = 3.5 kpc Gas Stars DM Total 9.50 log 10 M /M < 9.75 N gal = Vc(r) [km s 1 ] rc = 3.4 kpc 9.75 log 10 M /M < N gal = 513 rc = 3.3 kpc log 10 M /M < N gal = 463 rc = 3.2 kpc Reyes et al. (2011) log 10 M /M < N gal = rc = 3.1 kpc log 10 M /M < N gal = 222 rc = 3.0 kpc log M /M < log 10 M /M < N gal = 139 r [kpc] N gal = 60 rc = 2.9 kpc Vc(r) [km s 1 ] r [kpc] r [kpc] r [kpc] Figure 7. Simulated circular velocity curves and observed spiral galaxy rotation curves in different stellar mass bins. The green diamonds with error bars correspond to the total circular velocity and the RMS scatter around the mean. The black squares, red stars and blue circles represent the mean contributions of dark matter, star and gas particles respectively. The dashed vertical line is the conservative resolution limit, r c. The background brown curves are the best-fit Hα rotation curves extracted from Reyes et al. (2011). We plot their data up to their i band measured isophotal R 80 radii. follow the NFW prediction in the outer parts, but the inner profile is significantly steeper than the NFW form, which has an inner slope (ρ(r 0) = r η with η 1). The deviations from an NFW profile can be quite large on small scales. To show this, we fit the total mass profiles using the fitting procedure defined by Neto et al. (7). We fit an NFW profile to the stacked profiles over the radial range [0.05, 1]R vir, shown respectively as blue dashed curves and filled circles in Fig. 8. This choice of minimum radius is larger than the conservative convergence radius given by version of the Power et al. (3) criterion that we adopted in the previous section. As described in Section 4.2, the bins are spherical and spaced logarithmically in radius. The Neto et al. (7) fit is performed by minimizing a χ 2 expression with two free parameters, r s and δ c, characterising the NFW profile, over a set of N b (= 17) radial bins. We use the Levenberg & Marquart method to minimize the RMS deviation, σ fit, between the binned logarithmic densities ρ i and the NFW profile ρ NFW: N 1 b σ fit = (log N b 1 10 ρ i log 10 ρ NFW(δ c, r s)) 2. (16) i=1 Note that the bins are weighted equally. The best-fit profile for each stacked halo mass bin is shown in Fig. 8 as a blue dashed line. The NFW profile is a very good fit to the filled circles, confirming that the outer parts of the halos are well described by this profile within R. However, the NFW profile is clearly a poor fit at small radii (r 0.05R vir) for halos with a significant stellar mass, i.e. for halos above 3 1 M, as

12 12 M. Schaller et al. r 2 ρ(r)/(r 2 ρ cr) rc = 1.9 kpc rc = 1.8 kpc rc = 1.8 kpc rc = 1.8 kpc M = M M = M M = 4 0 M M = M r 2 ρ(r)/(r 2 ρ cr) rc = 3.9 kpc rc = 3.7 kpc rc = 3.6 kpc rc = 3.4 kpc M = 1 1 M M = M M = M M = 4 1 M r 2 ρ(r)/(r 2 ρ cr) rc = 3.3 kpc rc = 3.2 kpc rc = 3.1 kpc rc = 3.0 kpc M = M M = 1 2 M M = M M = M r 2 ρ(r)/(r 2 ρ cr) rc = 3.0 kpc rc = 2.9 kpc rc = 2.9 kpc rc = 2.9 kpc M = 4 2 M M = M M = 1 3 M M = M r 2 ρ(r)/(r 2 ρ cr) rc = 2.9 kpc rc = 2.8 kpc M = M M = 4 3 M M = M M = 1 4 M r/r r/r r/r r/r 1 Figure 8. Stacked density profiles of the total mass normalized by the average R radius and scaled by r 2 for halos of different masses. The filled circles are the data points used to fit an NFW profile following Neto et al. (7), i.e. radial bins above data points below it are shown using fainter symbols. The blue dashed lines correspond to the NFW fit to the filled circles, while the brown lines correspond to an Einasto profile fit to all radial bins down to the convergence radius, r c. The red solid line is the best-fit profile given by Eqn. 19, which includes an NFW contribution for the outer parts of the halos and an additional contribution around the centre to model the baryons. The best-fitting parameters for each mass bins are given in Table 2. expected from Fig. 6, due to the increased contribution of the stars and the subsequent contraction of the DM profile. For halo masses above 2 M, the discrepancy between the NFW prediction and the actual total mass density profile reaches factors of two close to the resolution limit. When multiplied by r 2, the NFW profile reaches a maximum at r = r s. For M > 3 1 M the profiles do not display a single sharp maximum but rather a broad range of radii at almost constant r 2 ρ(r), i.e. a quasi isothermal profile. For M 3 3 M, the difference is even more striking as a sec-

13 Baryon effects on the internal structure of ΛCDM halos 13 ond maximum appears at small radii. We will explore alternative fitting formula in what follow, but it is clear that a fitting formula describing the most massive halos will require several parameters to work well. In their detailed study, Navarro et al. (4) explored the use of a more general class of profiles, where the slope varies with radius as a power law. This alternative profile was originally introduced by Einasto (1965) to model old stellar populations in the Milky Way, and so Navarro et al. (4) called it the Einsasto profile : [ ρ(r) = ρ 2 exp 2 (( ) α r 1)], (17) α which can be rewritten as ( ) α d ln ρ(r) r = 2, (18) d ln r to highlight that the slope is a power-law of radius. Navarro et al. (4) showed that halos in simulations are typically better fit by the Einasto profile and that the value of the power law parameter, α 0.17, can be used across the whole simulated halo mass range. This was confirmed by Gao et al. (8) and Duffy et al. (8) who found a weak dependence of α on the peak-height parameter. Gao et al. (8) demonstrated that the Einasto profile is more robust to choices of the minimal converged radius, r c, improving the quality of the fit. In the case of our sample of halos, the additional freedom to change the slope of the power law describing the density profile helps improve the fit. We use the same procedure as in the NFW case to find the best-fitting parameters (, ρ 2, α) but instead of using only the radial bins with r > 0.05R vir, we use all bins with r > r c. The number of bins used is now a function of the halo mass. The resulting best-fit profiles are displayed in Fig. 8 as solid yellow lines. The fits are slightly better than in the NFW case simply because the rolling power law allows for a wider peak in r 2 ρ(r), but the Einasto profile is clearly unable to capture the complex behaviour seen in the profiles of the highest mass bins. The better fit quality is only incidental. Furthermore, if we had used the full range of radial bins for the NFW fitting procedure, we would have obtained similar fits as the two functions are very similar. Similarly, restricting the Einasto fit to the bins with r > 0.05R vir yields a best fit profile (and σ fit ) almost identical to the NFW ones shown by the dashed blue lines. Clearly, in the presence of baryons, neither the NFW nor the Einasto profile faithfully represents the inner matter density profile. As Fig. 6 showed, the inner profile is shaped by both a substantial stellar contribution and the contraction of the dark matter associated with the elevated baryon fraction towards the centre. We find that the total profile can be fit everywhere by the following formula: ρ(r) ρ cr = δ c (r/r s) (1 + r/r s) 2 + δ i (r/r ( i) 1 + (r/r 2). (19) i) The first term is the NFW profile, which we have shown gives a good fit to the outer, DM-dominated profile. The second term is NFW-like in that is shares the same asymptotic behaviour at small and large radii and has a slope of -2 at its scale radius, r = r i. We have found by trial and error that its sharper transition relative to the NFW profile between the asymptotic slope regimes of -1 and -3, which causes it to rise a factor of two above a corresponding NFW profile that shares the same scale radius and asymptotic behaviour at small and large radii, make it particularly suitable for describing the deviations in the density profiles above an NFW profile seen in the central regions of the EAGLE halos. We fit this profile using all the radial bins down to our resolution limit, r c. We rewrite expression (16) using our new profile and minimize σ fit leaving the four parameters (r s, δ c, r i, δ i) free. The resulting fits are displayed in Fig. 8 as red solid lines. The values of the best-fitting parameters are given in Table 2. The fit is clearly of a much better quality than the NFW and Einasto formulas for the same set of radial bins. For the lowest mass halos (M < 6 0 M ), this new profile does not provide a better σ fit than a standard NFW profile does. This is expected since the baryons have had little impact on their inner structure. The values of r i and δ i are, hence, not constrained by the fits. For these low mass stacks, we only provide the best-fitting NFW parameters in Table 2 instead of the parameters of our alternative profile. The different features of the simulated halos are well captured by the additional component of our profile. We will demonstrate in the next sections that the additional degrees of freedom can be recast as physically meaningful quantities and that these are closely correlated with the halo mass. As in the case of the NFW profile, this implies that this new profile is effectively a one parameter fit, where the values of all the four parameters depend solely on the mass of the halo. It is worth mentioning that this profile also reproduces the trends in the radial bins below the resolution limit r c. For completeness, we give the analytic expressions for both the enclosed mass, M(r < R), and the gravitational potential, Φ(r), for the empirical profile of Eqn. 19, M(r < R) = 2πρ cr 2δ crs 3 and Φ(r) = 4πGρ cr [ ) ln (1 + Rrs R R + r s ) +δ iri 3 ln (1 + R2, (20) δcr3 s r r 2 i ) ln (1 + rrs r 2 i ] (21) [ ( ) )] +δ iri 2 π r 2 arctan + (1 ri r i 2r ln + r2. The expressions for an NFW profile are recovered by setting δ i = 0. Finally, we stress that while this function provides an excellent fit to the results over the range of applicability the second term should not be interpreted as a description of the stellar profile. Rather, the second term models a combination of the effect of all components, including the contraction of the dark matter, and is only valid above our resolution limit which is well outside the stellar half-mass radius. Higher-resolution simulations, with improved subgrid models, would be needed to model accurately the stars and gas in these very inner regions. 4.5 Dark matter density profile It is interesting to see whether the radial distribution of dark matter is different in the and EAGLE simulations. In this subsection we look at the density profiles of just the DM in both the and EAGLE simulations. In Fig. 9 we show the profiles of the

14 14 M. Schaller et al. Table 2. Best-fit parameters for the profile (Eqn. 19) for each stack of relaxed halos as plotted in Fig. 8. The tabulated values correspond to the black circles plotted in Figs. 13, 14 and 15. The first column gives the centre of the mass bin used for each stack and the last column the number of halos in each of the stacks. The concentration, c, and inner profile mass, M i, are defined, respectively, by Eqns. 22 and 25. For the halo stacks in the lowest mass bins, the profile 19 does not provide a better fit than a standard NFW. We hence only give the best-fitting parameters to the NFW fit. M [M ] R [kpc] r s [kpc] c [ ] δ c [ ] r i [kpc] δ i [ ] M i [M ] N halo r 2 ρ(r)/(r 2 ρ cr) rc = 4.2 kpc rc = 3.6 kpc M = 1 1 M M = 1 2 M M = 1 3 M M = 1 4 M r/r r/r r/r r/r 1 rc = 3.2 kpc rc = 3.1 kpc Figure 9. Stacked density profiles of the halos normalized by the average R radius and scaled by r 2 for a selection of masses. The filled circles are the data points used to fit an NFW profile following Neto et al. (7). The vertical line shows the resolution limit. Data points are only shown at radii larger than the Plummer-equivalent softening (2.8ɛ = 0.7 kpc). The blue dashed and solid brown lines correspond, respectively, to the best-fit NFW and Einasto profiles to the filled circles. Only one halo contributes to the right hand panel. stacked halos extracted from the simulation for different halo mass bins. The dark matter outside 0.05R vir is well fit by the NFW profile, in agreement with previous work. The yellow curves show the best fit Einasto profile, and in agreement with many authors (Navarro et al. 4; Gao et al. 8; Dutton & Macciò 2014) we find that the Einasto fit, with one extra parameter, provides a significantly better fit to the inner profile. We show the stacked DM density profiles for the EAGLE simulation in Fig. 10 together with NFW and Einasto fits to the density at 0.05 r/r vir 1. For the radii beyond 0.05R vir the NFW profile provides a good fit. The Einasto profile fits are better in the inner regions, but for the middle two mass bins (2 M and 3 M ), the DM profile rises significantly above the Einasto fit. This rise coincides with a more pronounced feature in the total mass profile. The peak of the central stellar mass fraction occurs at this same halo mass scale, as shown in Fig. 4. We conclude that the DM components of our simulated halos in both the and EAGLE simulations are well described by an NFW profile for radii [0.05R R ]. For the simulation an Einasto profile provides a better fit than an NFW profile at smaller radii. However, for the EAGLE simulation neither an NFW nor the Einasto profile provide a particularly good fit inside 0.05R vir for halos in the 2 M and 3 M mass bins, where the contribution of stars to the inner profile is maximum. For less massive and more massive halos than this both functions give acceptable fits. In their detailed study of ten simulated galaxies from the MaGICC project (Stinson et al. 2013), Di Cintio et al. (2014) fitted (α, β, γ)-profiles (Jaffe 1983) to the DM profiles of haloes in the mass range 0 M M vir 2 M and studied the dependence of the parameters on the stellar fraction. We leave the study of the DM profiles in the EAGLE halos to future work but we

15 Baryon effects on the internal structure of ΛCDM halos 15 r 2 ρ(r)/(r 2 ρ cr) rc = 3.9 kpc rc = 3.2 kpc rc = 2.9 kpc M = 1 1 M M = 1 2 M M = 1 3 M M = 1 4 M r/r r/r r/r r/r 1 rc = 3.1 kpc Figure 10. Stacked density profiles of the dark matter component of the EAGLE halos normalized by the average R radius and scaled by r 2 for a selection of halo masses. The green dash dotted line represents the total mass profile (from Fig. 8. The vertical line shows the resolution limit. Data points are only shown at radii larger than the Plummer-equivalent softening (2.8ɛ = 0.7 kpc). The blue dashed lines and solid brown lines correspond, respectively, to the best-fit NFW and Einasto profiles to the filled circles. note that although in the small halo regime, M 2 M, an (α, β, γ)-profile may be a good fit, the profiles of our most massive halos, M 3 M, show varying slopes down to small radii, r 0.05R vir, and are unlikely to be well fit by such a function as was already suggested by Di Cintio et al. (2014). 4.6 Halo concentrations The concentration of a halo, c X, is conventionally defined by the ratio, c X = R X/r conc, where R X is the radius within which mean internal density is Xρ cr, and r conc is the radius at which the spherically averaged density profile (assumed monotonic) obeys d ln ρ(r) d ln r = 2. (22) For an NFW profile, r conc = r s, while for an Einasto profile r conc =. We set X =. Previous work (Navarro et al. 1997; Avila-Reese et al. 1999; Jing 0; Bullock et al. 1; Eke et al. 1; Zhao et al. 3; Neto et al. 7; Macciò et al. 7; Duffy et al. 8; Gao et al. 8; Dutton & Macciò 2014) has shown that the concentration and the mass of relaxed halos are anticorrelated (at z = 0), and follow a power law of the form ( ) B M c = A, (23) 4 h 1 M where A 5 and B 0.1. The best-fit values of these parameters are sensitive to the cosmological parameters, particularly to the values of σ 8 and Ω m (e.g. Duffy et al. 8; Dutton & Macciò 2014). The value of c at redshift zero is linked to the background density of the Universe at the time of formation of the halo (Navarro et al. 1997; Ludlow et al. 2013) which is affected by σ 8 and Ω m. Higher values of these parameters lead to earlier halo formation times at a given mass and therefore higher concentrations. The concentrations of individual halos of a given mass scatter about the median value with an approximately log-normal distribution (Jing 0; Neto et al. 7). The amplitude of this scatter decreases with halo mass (Neto et al. 7). While formally Eqn. 22 implicitly defines R conc, it is impractical to apply a differential measure of the density to determine the concentrations of individual halos, even in simulations, because the density profiles are noisy and sensitive to the presence of substructures. In practice, the concentration is determined by fitting the spherically averaged density profile over a range of radii encom- Table 3. Best fitting parameters and their 1σ uncertainty for the massconcentration relation (Eqn. 23) of the stacks of relaxed halos. The values correspond to those shown in the legends in Fig. 11. From top to bottom: NFW fit to the halos, NFW fit to the total mass of the EAGLE halos, and NFW fit to the dark matter component of the EAGLE halos. All profiles were fit over the radial range [0.05 1]R vir. The uncertainties are taken to be the diagonal elements of the correlation matrix of the least-squares fitting procedure. Fit A B c, 5.22 ± ± c,tot,nfw ± ± c,dm,nfw ± ± passing r s with a model. This approach only works if the model provides a good description of the true halo profile over the fitted range. We have shown in Section 4.4 that the density profiles of halos in both the EAGLE and simulations are well described by an NFW profile over the range [0.05 1]R vir, so we fit an NFW model over this range. Fig. 11 shows the NFW concentration of relaxed halos as a function of halo mass for the and EAGLE simulations. The top panel shows the simulation. The black line is the best fit power law of Eqn. 23 to the solid black circles (corresponding to the stacks containing at least five halos) using Poissonian errors for each bin. We have verified that fitting individual halos (faint green circles in the same figure) returns essentially the same values of A and B. Table 3 lists the best-fitting values of these parameters. It is worth mentioning that the best-fitting power laws fit the halo stacks in the simulations equally well. The mass-concentration relation of Dutton & Macciò (2014) is shown as a red dashed line in the top panel of Fig. 11. This fit is based on a series of cosmological simulations of a ΛCDM model very similar to ours with the cosmological parameters values taken from the Planck Collaboration et al. (2014) data. Using several volumes at different resolutions, they were able to determine the concentration-mass relation over the range 0 M < M < M at z = 0. Fitting an NFW model to estimate the concentration, as we do here, they obtained ( ) M c = 5.05, (24) 4 h 1 M which agrees well with our results.

16 16 M. Schaller et al. c,,nfw c = A ( M /4 h 1 M ) B A = ± B = ± z = M [M ] 4 5 r EAGLE /r s EAGLE s z = 0 c,tot,nfw c = A ( M /4 h 1 M ) B A = ± B = ± M, [M ] Figure 12. Ratio of NFW scale radii, r s, in matched relaxed halos in the and EAGLE simulations. The black points are placed at the geometric mean of the ratios in each mass bin. c,dm,nfw z = M [M ] c = A ( M /4 h 1 M ) B A = ± B = ± z = M [M ] 4 5 Figure 11. Halo concentration, c, as a function of mass M. The top panel shows the simulation fit with the canonical NFW profile over the range [0.05 1]R vir. The middle panel shows the same fit applied to the total matter density profiles of the EAGLE halos. The bottom panel shows the same fit to just the dark matter in the EAGLE halos. The faint coloured points in each panel are the values for individual halos and the black circles the values for the stacked profiles in each mass bin. Halos and stacks with M < 6 0 M are taken from the L025N0752 simulation whilst the higher mass objects have been extracted from the L100N1504 simulation. The solid black line is the best-fit power law (Eqn. 23) to the solid black circles. The best-fit parameters are shown in each panel. The best-fit power law to the halos is repeated in the other panels as a dashed line. The red dashed line on the first panel is the best-fit relation from Dutton & Macciò (2014). mass - concentration relation are given in the second line of Table 3. Both the amplitude and slope are consistent with the values for the simulation. As discussed in Section 3.1, matched halos in the and EAGLE simulations have, on average, a lower mass in the EAGLE simulation. For the smallest halos, the average ratio is as low as Because of this shift in mass, some difference in the concentration-mass relation might be expected between the two simulations but, since the value of the slope is small and , the effect on the amplitude is also small. A consequence of the shift in M is that the relative sizes of R for matched halos is R EAGLE /R 0.9. In Fig. 12 we show that the mean ratio of rs EAGLE /rs for matched relaxed halos is also slightly below unity, so the net effect of those two shifts is that the concentrations are very similar in both simulations. Finally, the bottom panel of Fig. 11 shows the concentration of the DM only component of EAGLE halos. We fit an NFW profile in the same way as for the total matter profiles in the panels above. As would be expected from the analysis of Fig. 8 and the fact that the outer parts of the dark halos are well described by the NFW profile, the same trend with mass can be seen as for the simulation. The best-fitting power law to the mass-concentration relation is given at the bottom of Table 3. The values of the parameters are again close to the ones obtained for both the EAGLE and the simulations. We stress that the agreement between the EAGLE and simulations breaks down if we include radii smaller than 0.05R vir in the fit. Hence, the mass - concentration relation given for EAGLE in Table 3 should only be used to infer the density profiles beyond 0.05R vir. Not unexpectedly, given the sensitivity of the concentration to changes in the cosmological parameters, the values for the fit we obtain for the simulation are significantly different from those reported by Neto et al. (7), Macciò et al. (7) and Duffy et al. (8). Compared to the latter, the slope (B) is steeper and the normalisation (A) is higher. This change can be attributed mainly to changes in the adopted cosmological parameters (σ 8, Ω m) which were (0.796, 0, 258) in Duffy et al. (8) and (0.8288, 0.307) here. The second panel of Fig. 11 shows the concentrations for the total matter density profiles of the EAGLE simulation obtained using the same fitting procedure. The best-fitting parameters for the 4.7 Best-fit parameter values for the new density profile We showed in Section 4.4 that the density profiles of halos in the EAGLE simulation are not well fit by an NFW profile in the inner regions, and we proposed Eqn. 19 as a new fitting formula for these profiles. This new profile has two lengthscales, r s and r i, where the former describes the NFW-like outer parts of the halo, and the latter the deviations from NFW in the inner regions. For lower-mass halos these two lengths become similar, so both terms of the profile can contribute significantly to the density at all radii. We can still define the concentration of a halo in this model as R /r s, but we would expect to obtain a different mass-concentration relation

17 Baryon effects on the internal structure of ΛCDM halos 17 c,tot,new c = A ( M /4 h 1 M ) B A = ± B = ± z = M [M ] 4 5 Figure 13. Halo concentration, c, as a function of mass, M, for the total matter density profiles of the EAGLE simulation using the fitting function of Eqn. 19 and the r s parameter to define the concentration, c = R /r s. The colour points are for individual halos and the black circles for the stacked profiles in each mass bin. The solid black line is the best-fit power law (Eqn. 23) to the solid black circles. The best-fit values are given in the legend at the top right. The dashed line shows the best fitting power law to the halos extracted from the simulation fitted using an NFW profile. ri [kpc] r i = 2.27 kpc ε = 700 pc 0.3 z = M [M ] 4 5 Figure 14. The characteristic radius, r i, of the central component as function of halo mass (Eqn. 19) for halos in the EAGLE simulation. The red squares correspond to all the halos fitted individually and the overlaying black circles to the stacked halos in each mass bin. Stacks containing less than three objects are shown as open circles. The minimum Plummerequivalent softening length (ɛ = 0.7 kpc) is indicated by the grey dashed line at the bottom of the figure. The average value of the stacks with more than three objects is indicated by a solid black line. from that for the dark matter-only case. Fig. 13 shows this relation for relaxed EAGLE halos. The anticorrelation seen when fitting an NFW profile is still present and we can use the same power-law formulation to describe the mass-concentration relation of our halo stacks. The values of the best-fit parameters, given in the figure, differ significantly from those obtained using the NFW fits listed in Table 3. We now consider the two remaining parameters of the profile described by Eqn. 19. The inner component is characterized by two quantities, a scale radius, r i, and a density contrast, δ i. We stress that this inner profile should not be interpreted as the true underlying model of the galaxy at the centre of the halo. It is an empirical model that describes the deviation from NFW due to the presence of stars and some contraction of the dark matter. The profiles have been fit using the procedure described in Section 4.4 using all radial bins with r > r c. Mi [M ] z = 0 Halo M M [M ] Figure 15. The mass, M i, defined in Eqn. 25, as a function of halo mass, M. The red squares correspond to the individual halos and the overlaying black circles to the stacked profiles. The green solid line is the stellar mass - halo mass relation from the EAGLE simulation (Schaye et al. 2015). The dependence of the r i scale radius on the halo mass is shown in Fig. 14. The radius r i is roughly constant over the entire halo mass range in the simulation. The scatter is large at all masses, but there is a weak trend with mass in the low-mass regime. This regime is, however, difficult to study as may be seen in the first few panels of Fig. 8: for the smallest halos, the effects due to baryons are small and the profile is thus closer to NFW than for the highermass bins. The empirical profile (Eqn. 19) tends towards an NFW profile as δ i 0 or r i 0. We find that, for the smallest halos, there is a degeneracy between these two parameters and the values of r i and δ i can be changed by an order of magnitude (self-consistently) without yielding a significantly different σ fit value. This is not a failure of the method but rather a sign that the baryonic effects on the profile shape become negligible for the lowest-mass halo, at least for the range of radii resolved in this study. Rather than working with the δ i and r i parameters, we can combine them into a single parameter that reflects the additional mass contained in the central parts of the halo above and above that from the NFW component. Integrating the inner profile up to r i, we can obtain an estimate of this additional mass which we define as: M i = (2π ln 2)ρ crr 3 i δ i 4.355ρ crr 3 i δ i. (25) If r i were really constant, then M i would simply be a proxy for δ i. The mass, M i, is shown in Fig. 15 as a function of the halo mass, M. The black points corresponding to the stacked profiles lie in the middle of the relation for individual halos. The mass, M i, increases with halo mass. For halos with M 2 M, the fraction, M i/m, increases with M highlighting that the effect of the baryons is more important for the bigger halos. This could have been expected by a careful inspection of Fig. 4, which shows that the central stellar and baryonic fractions peak at M 2 M. For larger halos, the M -M i relation flattens reflecting the decrease in stellar fractions seen at the centre of the largest EAGLE halos. To confirm this conjecture, we plot the stellar mass - halo mass relation for the EAGLE simulation as a solid green line in the same figure (Schaye et al. 2015) 1. Neglecting the two highest mass bins (open circles), the similarity between this relation and our somewhat arbitrary definition of M i seems to indicate that the stellar 1 Note that the EAGLE simulation reproduces abundance matching results (Schaye et al. 2015).

Joop Schaye (Leiden) (Yope Shea)

Joop Schaye (Leiden) (Yope Shea) Overview of sub-grid models in cosmological simulations Joop Schaye (Leiden) (Yope Shea) Length Scales (cm) Subgrid models Cosmological simulations 8 0 2 11 18 20 22 24 28 interparticle distance in stars

More information

Durham Research Online

Durham Research Online Durham Research Online Deposited in DRO: 16 February 2016 Version of attached le: Published Version Peer-review status of attached le: Peer-reviewed Citation for published item: Furlong, M. and Bower,

More information

Black Hole Feedback. What is it? What does it do? Richard Bower Durham University

Black Hole Feedback. What is it? What does it do? Richard Bower Durham University Black Hole Feedback What is it? What does it do? GALFORM: RGB + Benson, Lagos, Fanidakis, Frenk, Lacey, Baugh & Cole +++ EAGLE: Booth, Dalla Vecchia, Crain, Furlong, Rosas- Guevara, Schaye, RGB, Theuns,

More information

Dwarf Galaxies as Cosmological Probes

Dwarf Galaxies as Cosmological Probes Dwarf Galaxies as Cosmological Probes Julio F. Navarro The Ursa Minor dwarf spheroidal First Light First Light The Planck Satellite The Cosmological Paradigm The Clustering of Dark Matter The Millennium

More information

Towards Understanding Simulations of Galaxy Formation. Nigel Mitchell. On the Origin of Cores in Simulated Galaxy Clusters

Towards Understanding Simulations of Galaxy Formation. Nigel Mitchell. On the Origin of Cores in Simulated Galaxy Clusters Towards Understanding Simulations of Galaxy Formation Nigel Mitchell On the Origin of Cores in Simulated Galaxy Clusters Work published in the Monthly Notices of the Royal Astronomy Society Journal, 2009,

More information

The drop in the cosmic star formation rate below redshift 2 is caused by a change in the mode of gas accretion and by active galactic nucleus feedback

The drop in the cosmic star formation rate below redshift 2 is caused by a change in the mode of gas accretion and by active galactic nucleus feedback 3 The drop in the cosmic star formation rate below redshift 2 is caused by a change in the mode of gas accretion and by active galactic nucleus feedback The cosmic star formation rate is observed to drop

More information

The Los Cabos Lectures

The Los Cabos Lectures January 2009 The Los Cabos Lectures Dark Matter Halos: 2 Simon White Max Planck Institute for Astrophysics EPS statistics for the standard ΛCDM cosmology Millennium Simulation cosmology: Ωm = 0.25, ΩΛ

More information

The IGM in simulations

The IGM in simulations The IGM in simulations Institute for Computational Cosmology Ogden Centre for Fundamental Physics Durham University, UK and University of Antwerp Belgium 1 Menu: Simulated DLAs: column density and dynamics

More information

The separate and combined effects of baryon physics and neutrino free streaming on large-scale structure

The separate and combined effects of baryon physics and neutrino free streaming on large-scale structure Advance Access publication 2017 June 20 doi:10.1093/mnras/stx1469 The separate and combined effects of baryon physics and neutrino free streaming on large-scale structure Benjamin O. Mummery, 1 Ian G.

More information

arxiv: v1 [astro-ph.ga] 19 Oct 2015

arxiv: v1 [astro-ph.ga] 19 Oct 2015 Mon. Not. R. Astron. Soc. 000, 1 17 (2015) Printed 21 October 2015 (MN LATEX style file v2.2) Size evolution of normal and compact galaxies in the EAGLE simulation arxiv:1515645v1 [astro-ph.ga] 19 Oct

More information

THE BOLSHOI COSMOLOGICAL SIMULATIONS AND THEIR IMPLICATIONS

THE BOLSHOI COSMOLOGICAL SIMULATIONS AND THEIR IMPLICATIONS GALAXY FORMATION - Durham -18 July 2011 THE BOLSHOI COSMOLOGICAL SIMULATIONS AND THEIR IMPLICATIONS JOEL PRIMACK, UCSC ΛCDM Cosmological Parameters for Bolshoi and BigBolshoi Halo Mass Function is 10x

More information

Origin and Evolution of Disk Galaxy Scaling Relations

Origin and Evolution of Disk Galaxy Scaling Relations Origin and Evolution of Disk Galaxy Scaling Relations Aaron A. Dutton (CITA National Fellow, University of Victoria) Collaborators: Frank C. van den Bosch (Utah), Avishai Dekel (HU Jerusalem), + DEEP2

More information

arxiv: v2 [astro-ph.ga] 2 Oct 2014

arxiv: v2 [astro-ph.ga] 2 Oct 2014 Mon. Not. R. Astron. Soc. 000, 000 000 (0000) Printed 3 October 2014 (MN LATEX style file v2.2) The EAGLE project: Simulating the evolution and assembly of galaxies and their environments arxiv:1407.7040v2

More information

Feedback from massive stars in dwarf galaxy formation

Feedback from massive stars in dwarf galaxy formation Highlights of Spanish Astrophysics VII, Proceedings of the X Scientific Meeting of the Spanish Astronomical Society held on July 9-13, 2012, in Valencia, Spain. J. C. Guirado, L. M. Lara, V. Quilis, and

More information

The Iguaçu Lectures. Nonlinear Structure Formation: The growth of galaxies and larger scale structures

The Iguaçu Lectures. Nonlinear Structure Formation: The growth of galaxies and larger scale structures April 2006 The Iguaçu Lectures Nonlinear Structure Formation: The growth of galaxies and larger scale structures Simon White Max Planck Institute for Astrophysics z = 0 Dark Matter ROT EVOL Cluster structure

More information

Princeton December 2009 The fine-scale structure of dark matter halos

Princeton December 2009 The fine-scale structure of dark matter halos Princeton December 2009 The fine-scale structure of dark matter halos Simon White Max Planck Institute for Astrophysics The dark matter structure of CDM halos A rich galaxy cluster halo Springel et al

More information

arxiv: v4 [astro-ph.co] 11 Sep 2015

arxiv: v4 [astro-ph.co] 11 Sep 2015 Mon. Not. R. Astron. Soc. 000, 1 22 (2013) Printed 14 September 2015 (MN LATEX style file v2.2) The impact of angular momentum on black hole accretion rates in simulations of galaxy formation arxiv:1312.0598v4

More information

The mass of a halo. M. White

The mass of a halo. M. White A&A 367, 27 32 (2001) DOI: 10.1051/0004-6361:20000357 c ESO 2001 Astronomy & Astrophysics The mass of a halo M. White Harvard-Smithsonian Center for Astrophysics, Cambridge, MA 02138, USA e-mail: mwhite@cfa.harvard.edu

More information

The Dark Matter - Galaxy Connection: HOD Estimation from Large Volume Hydrodynamical Simulations

The Dark Matter - Galaxy Connection: HOD Estimation from Large Volume Hydrodynamical Simulations The Dark Matter - Galaxy Connection: HOD Estimation from Large Volume Hydrodynamical Simulations J. CASADO GÓMEZ (UAM) R. DOMÍNGUEZ-TENREIRO J. OÑORBE (UCA/Irvine) F. MARTINEZ - SERRANO (UMH) A. KNEBE

More information

Structure of Dark Matter Halos

Structure of Dark Matter Halos Structure of Dark Matter Halos Dark matter halos profiles: DM only: NFW vs. Einasto Halo concentration: evolution with time Dark matter halos profiles: Effects of baryons Adiabatic contraction Cusps and

More information

The enrichment history of cosmic metals

The enrichment history of cosmic metals CHAPTER 5 The enrichment history of cosmic metals Robert P. C. Wiersma, Joop Schaye, Claudio Dalla Vecchia, C. M. Booth, Tom Theuns, and Anthony Aguirre Accepted for publication in the Monthly Notices

More information

Cosmological simulations of galaxy formation

Cosmological simulations of galaxy formation Cosmological simulations of galaxy formation Modern galaxy formation simulations Mock gri SDSS composite image with dust absorption based on Draine opacity model. NGC4622 as seen from HST Outline - Impact

More information

2. What are the largest objects that could have formed so far? 3. How do the cosmological parameters influence structure formation?

2. What are the largest objects that could have formed so far? 3. How do the cosmological parameters influence structure formation? Einführung in die beobachtungsorientierte Kosmologie I / Introduction to observational Cosmology I LMU WS 2009/10 Rene Fassbender, MPE Tel: 30000-3319, rfassben@mpe.mpg.de 1. Cosmological Principles, Newtonian

More information

arxiv: v1 [gr-qc] 1 Dec 2017

arxiv: v1 [gr-qc] 1 Dec 2017 Can ΛCDM model reproduce MOND-like behavior? De-Chang Dai, Chunyu Lu Institute of Natural Sciences, Shanghai Key Lab for Particle Physics and Cosmology, School of Physics and Astronomy, Shanghai Jiao Tong

More information

The Formation and Evolution of Galaxy Clusters

The Formation and Evolution of Galaxy Clusters IAU Joint Discussion # 10 Sydney, July, 2003 The Formation and Evolution of Galaxy Clusters Simon D.M. White Max Planck Institute for Astrophysics The WMAP of the whole CMB sky Bennett et al 2003 > 105

More information

Effects of SN Feedback on the Dark Matter Distribution

Effects of SN Feedback on the Dark Matter Distribution The Galaxy Disk in Cosmological Context c 2008 International Astronomical Union Proceedings IAU Symposium No. IAU Symposium No.254, 2008 A.C. Editor, B.D. Editor & C.E. Editor, eds. Effects of SN Feedback

More information

Feedback and Galaxy Formation

Feedback and Galaxy Formation Heating and Cooling in Galaxies and Clusters Garching August 2006 Feedback and Galaxy Formation Simon White Max Planck Institute for Astrophysics Cluster assembly in ΛCDM Gao et al 2004 'Concordance'

More information

Formation and growth of galaxies in the young Universe: progress & challenges

Formation and growth of galaxies in the young Universe: progress & challenges Obergurgl. April 2014 Formation and growth of galaxies in the young Universe: progress & challenges Simon White Max Planck Institute for Astrophysics Ly α forest spectra and small-scale initial structure

More information

4. Structure of Dark Matter halos. Hence the halo mass, virial radius, and virial velocity are related by

4. Structure of Dark Matter halos. Hence the halo mass, virial radius, and virial velocity are related by 6-4-10see http://www.strw.leidenuniv.nl/ franx/college/galaxies10 10-c04-1 6-4-10see http://www.strw.leidenuniv.nl/ franx/college/galaxies10 10-c04-2 4. Structure of Dark Matter halos Obviously, we cannot

More information

Gaia Revue des Exigences préliminaires 1

Gaia Revue des Exigences préliminaires 1 Gaia Revue des Exigences préliminaires 1 Global top questions 1. Which stars form and have been formed where? - Star formation history of the inner disk - Location and number of spiral arms - Extent of

More information

arxiv:astro-ph/ v1 27 Nov 2000

arxiv:astro-ph/ v1 27 Nov 2000 A&A manuscript no. (will be inserted by hand later) Your thesaurus codes are: 02 (3.13.18) - methods: N-body simulations ASTRONOMY AND ASTROPHYSICS The mass of a halo Martin White arxiv:astro-ph/0011495v1

More information

The edge of darkness, and other halo surprises

The edge of darkness, and other halo surprises The edge of darkness, and other halo surprises Benedikt Diemer ITC Fellow, Harvard-Smithsonian Center for Astrophysics (in collaboration with Andrey Kravtsov and Surhud More) APEC Seminar IPMU 4/7/26 Strength

More information

The dark matter crisis

The dark matter crisis The dark matter crisis Ben Moore Department of Physics, Durham University, UK. arxiv:astro-ph/0103100 v2 8 Mar 2001 Abstract I explore several possible solutions to the missing satellites problem that

More information

Recent Progress in Modeling of Galaxy Formation. Oleg Gnedin (University of Michigan)

Recent Progress in Modeling of Galaxy Formation. Oleg Gnedin (University of Michigan) Recent Progress in Modeling of Galaxy Formation Oleg Gnedin (University of Michigan) In current simulations, galaxies look like this: 10 kpc Disk galaxy at z=3: stars, molecular gas, atomic gas (Zemp,

More information

The Formation of Galaxies: connecting theory to data

The Formation of Galaxies: connecting theory to data Venice, October 2003 The Formation of Galaxies: connecting theory to data Simon D.M. White Max Planck Institute for Astrophysics The Emergence of the Cosmic Initial Conditions > 105 independent ~ 5 measurements

More information

Durham Research Online

Durham Research Online Durham Research Online Deposited in DRO: 12 October 2017 Version of attached le: Published Version Peer-review status of attached le: Peer-reviewed Citation for published item: Barnes, David J. and Kay,

More information

4. Structure of Dark Matter halos. Hence the halo mass, virial radius, and virial velocity are related by

4. Structure of Dark Matter halos. Hence the halo mass, virial radius, and virial velocity are related by 13-4-12see http://www.strw.leidenuniv.nl/ franx/college/galaxies12 12-c04-1 13-4-12see http://www.strw.leidenuniv.nl/ franx/college/galaxies12 12-c04-2 4. Structure of Dark Matter halos Obviously, we cannot

More information

How feedback shapes the galaxy stellar mass function

How feedback shapes the galaxy stellar mass function How feedback shapes the galaxy stellar mass function Eagle, owls and other Gimics Institute for Computational Cosmology Ogden Centre for Fundamental Physics Durham University, UK and University of Antwerp

More information

Some useful spherically symmetric profiles used to model galaxies

Some useful spherically symmetric profiles used to model galaxies Some useful spherically symmetric profiles used to model galaxies de Vaucouleurs R 1/4 law for ellipticals and bulges A good fit to the light profile of many ellipticals and bulges: (constant such that

More information

N-body Simulations. Initial conditions: What kind of Dark Matter? How much Dark Matter? Initial density fluctuations P(k) GRAVITY

N-body Simulations. Initial conditions: What kind of Dark Matter? How much Dark Matter? Initial density fluctuations P(k) GRAVITY N-body Simulations N-body Simulations N-body Simulations Initial conditions: What kind of Dark Matter? How much Dark Matter? Initial density fluctuations P(k) GRAVITY Final distribution of dark matter.

More information

Galaxies 626. Lecture 3: From the CMBR to the first star

Galaxies 626. Lecture 3: From the CMBR to the first star Galaxies 626 Lecture 3: From the CMBR to the first star Galaxies 626 Firstly, some very brief cosmology for background and notation: Summary: Foundations of Cosmology 1. Universe is homogenous and isotropic

More information

arxiv:astro-ph/ v1 19 Nov 1999

arxiv:astro-ph/ v1 19 Nov 1999 Where are the First Stars now? Simon D.M. White & Volker Springel Max-Planck-Institute for Astrophysics, Garching bei München, Germany arxiv:astro-ph/9911378v1 19 Nov 1999 Abstract. We use high-resolution

More information

Impact of baryon physics on dark matter structures: a detailed simulation study of halo density profiles

Impact of baryon physics on dark matter structures: a detailed simulation study of halo density profiles Mon. Not. R. Astron. Soc. 405, 2161 2178 (2010) doi:10.1111/j.1365-2966.2010.16613.x Impact of baryon physics on dark matter structures: a detailed simulation study of halo density profiles Alan R. Duffy,

More information

AGN feedback and its influence on massive galaxy evolution

AGN feedback and its influence on massive galaxy evolution AGN feedback and its influence on massive galaxy evolution Darren Croton (University of California Berkeley) Simon White, Volker Springel, et al. (MPA) DEEP2 & AEGIS collaborations (Berkeley & everywhere

More information

Hunting for dark matter in the forest (astrophysical constraints on warm dark matter)

Hunting for dark matter in the forest (astrophysical constraints on warm dark matter) Hunting for dark matter in the forest (astrophysical constraints on warm dark matter) ICC, Durham! with the Eagle collaboration: J Schaye (Leiden), R Crain (Liverpool), R Bower, C Frenk, & M Schaller (ICC)

More information

Angular Momentum Problems in Disk Formation

Angular Momentum Problems in Disk Formation Angular Momentum Problems in Disk Formation MPIA Theory Group Seminar, 07/03/2006 The Standard Picture Disks galaxies are systems in centrifugal equilibrium Structure of disks is governed by angular momentum

More information

Numerical Cosmology & Galaxy Formation

Numerical Cosmology & Galaxy Formation Numerical Cosmology & Galaxy Formation Lecture 13: Example simulations Isolated galaxies, mergers & zooms Benjamin Moster 1 Outline of the lecture course Lecture 1: Motivation & Historical Overview Lecture

More information

The EAGLE simulations of galaxy formation: calibration of subgrid physics and model variations

The EAGLE simulations of galaxy formation: calibration of subgrid physics and model variations Mon. Not. R. Astron. Soc. 000, 1 26 (2011) Printed 6 April 2015 (MN LATEX style file v2.2) The EAGLE simulations of galaxy formation: calibration of subgrid physics and model variations The formation and

More information

Disk Formation and the Angular Momentum Problem. Presented by: Michael Solway

Disk Formation and the Angular Momentum Problem. Presented by: Michael Solway Disk Formation and the Angular Momentum Problem Presented by: Michael Solway Papers 1. Vitvitska, M. et al. 2002, The origin of angular momentum in dark matter halos, ApJ 581: 799-809 2. D Onghia, E. 2008,

More information

The EAGLE simulations of galaxy formation: the importance of the hydrodynamics scheme

The EAGLE simulations of galaxy formation: the importance of the hydrodynamics scheme doi:10.1093/mnras/stv2169 The EAGLE simulations of galaxy formation: the importance of the hydrodynamics scheme Matthieu Schaller, 1 Claudio Dalla Vecchia, 2,3 Joop Schaye, 4 Richard G. Bower, 1 Tom Theuns,

More information

The cosmic distance scale

The cosmic distance scale The cosmic distance scale Distance information is often crucial to understand the physics of astrophysical objects. This requires knowing the basic properties of such an object, like its size, its environment,

More information

Baryon Census in Hydrodynamical Simulations of Galaxy Clusters

Baryon Census in Hydrodynamical Simulations of Galaxy Clusters Baryon Census in Hydrodynamical Simulations of Galaxy Clusters Susana Planelles (Univ.Trieste-INAF) Collaborators: S.Borgani (Univ. Trieste-INAF), G.Murante (INAF Torino), L.Tornatore (Univ. Trieste),

More information

The Evolution of Galaxy Angular Momentum

The Evolution of Galaxy Angular Momentum icc.dur.ac.uk/eagle KMOS redshift one spectroscopic survey The Evolution of Galaxy Angular Momentum The KROSS team: Harrison, Johnson, Tiley, Stott, Swinbank, Bower, Bureau, Smail, Bunker, Cirasuolo, Sobral,

More information

Comparing l-galaxies, galform and eagle

Comparing l-galaxies, galform and eagle V. Gonzalez-Perez /9 Comparing l-galaxies, galform and eagle Violeta Gonzalez-Perez @violegp Quan Guo (Postdam), Qi Guo (Beijing), Matthieu Schaller (Durham), Michelle Furlong (Durham), Richard Bower (Durham),

More information

Cosmic Structure Formation on Supercomputers (and laptops)

Cosmic Structure Formation on Supercomputers (and laptops) Cosmic Structure Formation on Supercomputers (and laptops) Lecture 4: Smoothed particle hydrodynamics and baryonic sub-grid models Benjamin Moster! Ewald Puchwein 1 Outline of the lecture course Lecture

More information

Rotation rates, sizes and star formation efficiencies of a representative population of simulated disc galaxies

Rotation rates, sizes and star formation efficiencies of a representative population of simulated disc galaxies Mon. Not. R. Astron. Soc. 427, 379 392 (2012) doi:10.1111/j.1365-2966.2012.21951.x Rotation rates, sizes and star formation efficiencies of a representative population of simulated disc galaxies I. G.

More information

Theoretical ideas About Galaxy Wide Star Formation! Star Formation Efficiency!

Theoretical ideas About Galaxy Wide Star Formation! Star Formation Efficiency! Theoretical ideas About Galaxy Wide Star Formation Theoretical predictions are that galaxy formation is most efficient near a mass of 10 12 M based on analyses of supernova feedback and gas cooling times

More information

Superbubble Feedback in Galaxy Formation

Superbubble Feedback in Galaxy Formation Superbubble Feedback in Galaxy Formation Ben Keller (McMaster University) James Wadsley, Samantha Benincasa, Hugh Couchman Paper: astro-ph/1405.2625 (Accepted MNRAS) Keller, Wadsley, Benincasa & Couchman

More information

Structure and substructure in dark matter halos

Structure and substructure in dark matter halos Satellites and Tidal Streams ING IAC joint Conference La Palma, May 2003 Structure and substructure in dark matter halos Simon D.M. White Max Planck Institute for Astrophysics 500 kpc A CDM Milky Way Does

More information

Cooling, dynamics and fragmentation of massive gas clouds: clues to the masses and radii of galaxies and clusters

Cooling, dynamics and fragmentation of massive gas clouds: clues to the masses and radii of galaxies and clusters of massive gas and radii of M. Rees, J. Ostriker 1977 March 5, 2009 Talk contents: The global picture The relevant theory Implications of the theory Conclusions The global picture Galaxies and have characteristic

More information

The distribution of neutral hydrogen around high-redshift galaxies and quasars in the EAGLE simulation

The distribution of neutral hydrogen around high-redshift galaxies and quasars in the EAGLE simulation doi:10.1093/mnras/stv1414 The distribution of neutral hydrogen around high-redshift galaxies and quasars in the EAGLE simulation Alireza Rahmati, 1,2 Joop Schaye, 3 Richard G. Bower, 4 Robert A. Crain,

More information

The Los Cabos Lectures

The Los Cabos Lectures January 2009 The Los Cabos Lectures Dark Matter Halos: 3 Simon White Max Planck Institute for Astrophysics Shapes of halo equidensity surfaces Group Jing & Suto 2002 Galaxy δ 100 2500 6250 Shapes of halo

More information

An off-center density peak in the Milky Way's Dark Matter halo?

An off-center density peak in the Milky Way's Dark Matter halo? An off-center density peak in the Milky Way's Dark Matter halo? AKA: Impact of baryonic physics on DM detection experiments Michael Kuhlen, Berkeley with Javiera Guedes, Annalisa Pillepich, Piero Madau,

More information

The Millennium Simulation: cosmic evolution in a supercomputer. Simon White Max Planck Institute for Astrophysics

The Millennium Simulation: cosmic evolution in a supercomputer. Simon White Max Planck Institute for Astrophysics The Millennium Simulation: cosmic evolution in a supercomputer Simon White Max Planck Institute for Astrophysics The COBE satellite (1989-1993) Two instruments made maps of the whole sky in microwaves

More information

X-ray and Sunyaev-Zel dovich Effect cluster scaling relations: numerical simulations vs. observations

X-ray and Sunyaev-Zel dovich Effect cluster scaling relations: numerical simulations vs. observations X-ray and Sunyaev-Zel dovich Effect cluster scaling relations: numerical simulations vs. observations Daisuke Nagai Theoretical Astrophysics, California Institute of Technology, Mail Code 130-33, Pasadena,

More information

Physics of Galaxies 2016 Exercises with solutions batch I

Physics of Galaxies 2016 Exercises with solutions batch I Physics of Galaxies 2016 Exercises with solutions batch I 1. Distance and brightness at low redshift You discover an interesting galaxy in the local Universe and measure its redshift to be z 0.053 and

More information

The first stars and primordial IMBHs

The first stars and primordial IMBHs The first stars and primordial IMBHs Ab initio predictions of black hole merger rates by the time LISA flies? Tom Abel Penn State Initial Conditions Time Evolution z=100 z=24 z=20.4 10 comoving kpc Cosmological

More information

Cosmological simulations of our Universe Debora Sijacki IoA & KICC Cambridge

Cosmological simulations of our Universe Debora Sijacki IoA & KICC Cambridge Cosmological simulations of our Universe Debora Sijacki IoA & KICC Cambridge Cosmo 17 Paris August 30th 2017 Cosmological simulations of galaxy and structure formation 2 Cosmological simulations of galaxy

More information

The formation and evolution of globular cluster systems. Joel Pfeffer, Nate Bastian (Liverpool, LJMU)

The formation and evolution of globular cluster systems. Joel Pfeffer, Nate Bastian (Liverpool, LJMU) The formation and evolution of globular cluster systems Joel Pfeffer, Nate Bastian (Liverpool, LJMU) Introduction to stellar clusters Open clusters: few - 10 4 M few Myr - few Gyr solar metallicity disk

More information

Dependence of the inner dark matter profile on the halo mass

Dependence of the inner dark matter profile on the halo mass Mon. Not. R. Astron. Soc. 344, 1237 1249 (2003) Dependence of the inner dark matter profile on the halo mass Massimo Ricotti Institute of Astronomy, Madingley Road, Cambridge CB3 0HA Accepted 2003 June

More information

AST541 Lecture Notes: Galaxy Formation Dec, 2016

AST541 Lecture Notes: Galaxy Formation Dec, 2016 AST541 Lecture Notes: Galaxy Formation Dec, 2016 GalaxyFormation 1 The final topic is galaxy evolution. This is where galaxy meets cosmology. I will argue that while galaxy formation have to be understood

More information

ASTR 610 Theory of Galaxy Formation Lecture 18: Disk Galaxies

ASTR 610 Theory of Galaxy Formation Lecture 18: Disk Galaxies ASTR 610 Theory of Galaxy Formation Lecture 18: Disk Galaxies Frank van den Bosch Yale University, spring 2017 The Structure & Formation of Disk Galaxies In this lecture we discuss the structure and formation

More information

The theoretical view of high-z Clusters. Nelson Padilla, PUC, Chile Pucón, November 2009

The theoretical view of high-z Clusters. Nelson Padilla, PUC, Chile Pucón, November 2009 The theoretical view of high-z Clusters Nelson Padilla, PUC, Chile Pucón, November 2009 The Plan: 1) To see what the observations are telling us using models that agree with the cosmology, and with other

More information

The contributions of matter inside and outside of haloes to the matter power spectrum

The contributions of matter inside and outside of haloes to the matter power spectrum 5 The contributions of matter inside and outside of haloes to the matter power spectrum Halo-based models have been very successful in predicting the clustering of matter. However, the validity of the

More information

Global Scaling Relations of Spiral Galaxies

Global Scaling Relations of Spiral Galaxies Global Scaling Relations of Spiral Galaxies Aaron A. Dutton Max Planck Institute for Astronomy (MPIA), Heidelberg, Germany IAU 311, Galaxy Masses as constraints to Formation Models, Oxford, July 2014 Outline

More information

Galaxy Formation! Lecture Seven: Galaxy Formation! Cosmic History. Big Bang! time! present! ...fluctuations to galaxies!

Galaxy Formation! Lecture Seven: Galaxy Formation! Cosmic History. Big Bang! time! present! ...fluctuations to galaxies! Galaxy Formation Lecture Seven: Why is the universe populated by galaxies, rather than a uniform sea of stars? Galaxy Formation...fluctuations to galaxies Why are most stars in galaxies with luminosities

More information

A5682: Introduction to Cosmology Course Notes. 11. CMB Anisotropy

A5682: Introduction to Cosmology Course Notes. 11. CMB Anisotropy Reading: Chapter 8, sections 8.4 and 8.5 11. CMB Anisotropy Gravitational instability and structure formation Today s universe shows structure on scales from individual galaxies to galaxy groups and clusters

More information

The EAGLE simulations of galaxy formation: calibration of subgrid physics and model variations

The EAGLE simulations of galaxy formation: calibration of subgrid physics and model variations doi:10.1093/mnras/stv725 The EAGLE simulations of galaxy formation: calibration of subgrid physics and model variations Robert A. Crain, 1,2 Joop Schaye, 1 Richard G. Bower, 3 Michelle Furlong, 3 Matthieu

More information

Implementing sub-grid treatments of galactic outflows into cosmological simulations. Hugo Martel Université Laval

Implementing sub-grid treatments of galactic outflows into cosmological simulations. Hugo Martel Université Laval Implementing sub-grid treatments of galactic outflows into cosmological simulations Hugo Martel Université Laval Leiden, June 19, 2013 GALACTIC OUTFLOWS Optical image of galaxy (Hubble Space Telescope)

More information

Moving mesh cosmology: The hydrodynamics of galaxy formation

Moving mesh cosmology: The hydrodynamics of galaxy formation Moving mesh cosmology: The hydrodynamics of galaxy formation arxiv:1109.3468 Debora Sijacki, Hubble Fellow, ITC together with: Mark Vogelsberger, Dusan Keres, Paul Torrey Shy Genel, Dylan Nelson Volker

More information

Visible Matter. References: Ryden, Introduction to Cosmology - Par. 8.1 Liddle, Introduction to Modern Cosmology - Par. 9.1

Visible Matter. References: Ryden, Introduction to Cosmology - Par. 8.1 Liddle, Introduction to Modern Cosmology - Par. 9.1 COSMOLOGY PHYS 30392 DENSITY OF THE UNIVERSE Part I Giampaolo Pisano - Jodrell Bank Centre for Astrophysics The University of Manchester - March 2013 http://www.jb.man.ac.uk/~gp/ giampaolo.pisano@manchester.ac.uk

More information

Structure formation in the concordance cosmology

Structure formation in the concordance cosmology Structure formation in the Universe, Chamonix, May 2007 Structure formation in the concordance cosmology Simon White Max Planck Institute for Astrophysics WMAP3 team WMAP3 team WMAP3 team WMAP3 team In

More information

arxiv: v2 [astro-ph.co] 23 Nov 2010

arxiv: v2 [astro-ph.co] 23 Nov 2010 Mon. Not. R. Astron. Soc. 000, 1 21 (2010) Printed 24 November 2010 (MN LATEX style file v2.2) Gas expulsion by quasar-driven winds as a solution to the over-cooling problem in galaxy groups and clusters

More information

SUPPLEMENTARY INFORMATION

SUPPLEMENTARY INFORMATION doi:10.1038/nature10445 Measuring gravitational redshift of galaxies in clusters The gravitational potential depth in typical galaxy clusters, expressed in terms of the velocity shift, is estimated at

More information

8.1 Structure Formation: Introduction and the Growth of Density Perturbations

8.1 Structure Formation: Introduction and the Growth of Density Perturbations 8.1 Structure Formation: Introduction and the Growth of Density Perturbations 1 Structure Formation and Evolution From this (Δρ/ρ ~ 10-6 ) to this (Δρ/ρ ~ 10 +2 ) to this (Δρ/ρ ~ 10 +6 ) 2 Origin of Structure

More information

Simulating the Universe

Simulating the Universe Simulating the Universe Christine Corbett Moran, Irshad Mohammed, Manuel Rabold, Davide Martizzi, Doug Potter, Aurel Schneider Oliver Hahn, Ben Moore, Joachim Stadel Outline - N-body codes: where do we

More information

Modelling the galaxy population

Modelling the galaxy population Modelling the galaxy population Simon White Max Planck Institut für Astrophysik IAU 277 Ouagadougou 1 The standard model reproduces -- the linear initial conditions -- IGM structure during galaxy formation

More information

The Current Status of Too Big To Fail problem! based on Warm Dark Matter cosmology

The Current Status of Too Big To Fail problem! based on Warm Dark Matter cosmology The Current Status of Too Big To Fail problem! based on Warm Dark Matter cosmology 172th Astronomical Seminar Dec.3 2013 Chiba Lab.M2 Yusuke Komuro Key Word s Too Big To Fail TBTF Cold Dark Matter CDM

More information

MASSACHUSETTS INSTITUTE OF TECHNOLOGY Physics Department Earth, Atmospheric, and Planetary Sciences Department. Final Exam

MASSACHUSETTS INSTITUTE OF TECHNOLOGY Physics Department Earth, Atmospheric, and Planetary Sciences Department. Final Exam MASSACHUSETTS INSTITUTE OF TECHNOLOGY Physics Department Earth, Atmospheric, and Planetary Sciences Department Physics 8.282J EAPS 12.402J May 20, 2005 Final Exam Name Last First (please print) 1. Do any

More information

The fine-scale structure of dark matter halos

The fine-scale structure of dark matter halos COSMO11, Porto, August 2011 The fine-scale structure of dark matter halos Simon White Max-Planck-Institute for Astrophysics COSMO11, Porto, August 2011 Mark Vogelsberger The fine-scale structure of dark

More information

Massive black hole formation in cosmological simulations

Massive black hole formation in cosmological simulations Institut d Astrophysique de Paris IAP - France Massive black hole formation in cosmological simulations Mélanie HABOUZIT Marta Volonteri In collaboration with Yohan Dubois Muhammed Latif Outline Project:

More information

Killing Dwarfs with Hot Pancakes. Frank C. van den Bosch (MPIA) with Houjun Mo, Xiaohu Yang & Neal Katz

Killing Dwarfs with Hot Pancakes. Frank C. van den Bosch (MPIA) with Houjun Mo, Xiaohu Yang & Neal Katz Killing Dwarfs with Hot Pancakes Frank C. van den Bosch (MPIA) with Houjun Mo, Xiaohu Yang & Neal Katz The Paradigm... SN feedback AGN feedback The halo mass function is much steeper than luminosity function

More information

Constraints on dark matter annihilation cross section with the Fornax cluster

Constraints on dark matter annihilation cross section with the Fornax cluster DM Workshop@UT Austin May 7, 2012 Constraints on dark matter annihilation cross section with the Fornax cluster Shin ichiro Ando University of Amsterdam Ando & Nagai, arxiv:1201.0753 [astro-ph.he] Galaxy

More information

high density low density Rayleigh-Taylor Test: High density medium starts on top of low density medium and they mix (oil+vinegar) Springel (2010)

high density low density Rayleigh-Taylor Test: High density medium starts on top of low density medium and they mix (oil+vinegar) Springel (2010) GAS MIXES high density Springel (2010) low density Rayleigh-Taylor Test: High density medium starts on top of low density medium and they mix (oil+vinegar) HOT HALO highest resolved density nth= 50x10

More information

arxiv: v3 [astro-ph.ga] 11 Jun 2010

arxiv: v3 [astro-ph.ga] 11 Jun 2010 Mon. Not. R. Astron. Soc. 000, 000 000 (0000) Printed 31 July 2018 (MN LATEX style file v2.2) An observer s view of simulated galaxies: disc-to-total ratios, bars, and (pseudo-)bulges arxiv:1001.4890v3

More information

Moment of beginning of space-time about 13.7 billion years ago. The time at which all the material and energy in the expanding Universe was coincident

Moment of beginning of space-time about 13.7 billion years ago. The time at which all the material and energy in the expanding Universe was coincident Big Bang Moment of beginning of space-time about 13.7 billion years ago The time at which all the material and energy in the expanding Universe was coincident Only moment in the history of the Universe

More information

Galaxies in dark matter halos: luminosity-velocity relation, abundance and baryon content

Galaxies in dark matter halos: luminosity-velocity relation, abundance and baryon content Galaxies in dark matter halos: luminosity-velocity relation, abundance and baryon content arxiv:1005.1289 arxiv:1002.3660 S. Trujillo-Gomez (NMSU) in collaboration with: A. Klypin (NMSU), J. Primack (UCSC)

More information

Clusters: Observations

Clusters: Observations Clusters: Observations Last time we talked about some of the context of clusters, and why observations of them have importance to cosmological issues. Some of the reasons why clusters are useful probes

More information

1.1 Large-scale properties of the Universe

1.1 Large-scale properties of the Universe 1 Our understanding of both the large-scale properties of our Universe and the processes through which galaxies form and evolve has greatly improved over the last few decades, thanks in part to new observational

More information

The Milky Way in the cosmological context. Andrey Kravtsov The University of Chicago

The Milky Way in the cosmological context. Andrey Kravtsov The University of Chicago The Milky Way in the cosmological context Andrey Kravtsov The University of Chicago Milky Way and Its Stars, KITP, 2 February 2015 Cosmological context: hierarchical structure formation from a Gaussian

More information